id
stringlengths 27
33
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 13
1.81M
|
---|---|---|---|
warning/0003/math-ph0003001.html | ar5iv | text | # Renormalization of the Regularized Relativistic Electron-Positron Field
## 1. Introduction
In relativistic quantum electrodynamics (QED) the quantized electron-positron field $`\mathrm{\Psi }(x)`$, which is an operator-valued spinor, is written formally as
(1)
$$\mathrm{\Psi }(x):=a(x)+b^{}(x),$$
where $`a(x)`$ annihilates an electron at $`x`$ and $`b^{}(x)`$ creates a positron at $`x`$. (We use the notation that $`x`$ denotes a space-spin point, namely $`x=(๐ฑ,\sigma )\mathrm{\Gamma }=^3\times \{1,2,3,4\}`$ and $`d^3x`$ denotes integration over $`^3`$ and a summation over the spin index.) More precisely, we take the Hilbert space, $``$ of $`L^2(^3)`$ spinors, i.e., $`=L^2(^3)^4`$. To specify the one-electron space we choose a subspace, $`_+`$ of $``$ and denote the orthogonal projector onto this subspace by $`P_+`$. The one-positron space is the (anti-unitary) charge conjugate, $`C_{}`$, of the orthogonal complement, $`_{}`$ with projector $`P_{}`$. In position space, $`(C(\psi ))(๐ฑ):=i\beta \alpha _2\overline{\psi (๐ฑ)}`$, whereas in momentum space $`(C(\widehat{\psi }))(๐ฉ):=i\beta \alpha _2\overline{\widehat{\psi }(๐ฉ)}`$. For later purposes we note explicitly that the Hilbert space $``$ can be written as the orthogonal sum
(2)
$$=_+_{}.$$
If $`f_\nu `$ is an orthonormal basis for $`_+`$ with $`\nu 0`$ and an orthonormal basis for $`_{}`$ with $`\nu <0`$ then $`a(x):=_{\nu =0}^{\mathrm{}}a(f_\nu )f_\nu (x)`$, and $`b(x):=_{\nu =\mathrm{}}^1b(f_\nu )\overline{f_\nu (x)}`$, where $`a^{}(f)`$ creates an electron in the state $`P_+f`$, etc. (For further details about the notation see Thaller , and also Helffer and Siedentop and Bach et al .
For free electrons and positrons, $`_+`$ and $`_{}`$ are, respectively, the positive and negative energy solutions of the free Dirac operator
(3)
$$D_0=๐ถ๐ฉ+m_0\beta ,$$
in which $`๐ถ`$ and $`\beta `$ denote the four $`4\times 4`$ Dirac matrices and $`๐ฉ=i`$. The number $`m_0`$ is the bare mass of the electron/positron. Perturbation theory is defined in terms of this splitting of $``$ into $`_+`$ and $`_{}`$. The electron/positron lines in Feynman diagrams are the resolvents of $`D_0`$ split in this way. As we shall see, this splitting may not be the best choice, ultimately, and another choice (which will require a fine-structure dependent, inequivalent representation of the canonical anticommutation relations) might be more useful.
The Hamiltonian for the free electron-positron field is
$$_0=d^3x:\mathrm{\Psi }(x)^{}D_0\mathrm{\Psi }(x):$$
with this choice of $`_+`$. The symbol $`::`$ denotes normal ordering, i.e., anti-commuting all $`a^{},b^{}`$ operators to the left (but ignoring the anti-commutators).
In this paper we investigate the effect of the Coulomb interaction among the particles, which is a quartic form in the $`a^\mathrm{\#},b^\mathrm{\#}`$ operators. ($`\mathrm{\#}`$ denotes either a star or no star.) The normal ordering will give rise to extra constants and certain quadratic terms. Our main point will be that by making an appropriate choice of $`_+`$ that is different from the usual one mentioned above, we can absorb the additional quadratic terms into a mass, wave function, and charge renormalization. This choice of $`_+`$ has to be made self consistently and, as we show, can be successfully carried out if some combination of the fine structure constant $`\alpha `$ (=1/137 in nature) and the ultraviolet cutoff $`\mathrm{\Lambda }`$ is small enough (but โsmallโ is actually โhugeโ since the condition is, roughly, $`\alpha \mathrm{log}\mathrm{\Lambda }<1`$).
The physical significance of our construction is not immediate, but it does show that certain annoying terms, when treated non-perturbatively, can be incorporated into the renormalization program. It is not at all clear that what we have called wave function renormalization, for example, really corresponds to a correct interpretation of that renormalization, or even what the true meaning of wave function renormalization is. It could also be interpreted as a renormalization of Planckโs constant. Likewise, the charge renormalization we find is not mandatory. But our main point is that the effect is mathematically real and is best dealt with by a change of the meaning of the one-particle electron/positron states. It should be taken into account in the non-perturbative QED that is yet to be born. Of course, there are other renormalization effects in QED that we do not consider. In particular, the magnetic field has not been included and so we do not have a Ward identity to help us. Our result here is only a small part of the bigger picture of a proper, non-perturbative QED.
The last section of this paper contains a brief discussion of some possible interpretations of our findings and the interested reader is urged to look at that section.
Our starting point is the unrenormalized (โbareโ) Hamiltonian for a quantized electron/positron field interacting with itself via Coulombโs law, namely
(4)
$$\begin{array}{c}^{\mathrm{bare}}\hfill \\ \hfill :=d^3x:\mathrm{\Psi }^{}(x)D_0\mathrm{\Psi }(x):+\frac{\alpha }{2}d^3xd^3yW(๐ฑ,๐ฒ):\mathrm{\Psi }^{}(x)\mathrm{\Psi }(x)::\mathrm{\Psi }^{}(y)\mathrm{\Psi }(y):.\end{array}$$
where $`W`$ is a symmetric ($`W(x,y)=W(y,x)`$) interaction potential. The case of interest here is the Coulomb potential $`W(x,y)=\delta _{\sigma ,\tau }|๐ฑ๐ฒ|^1`$ and regularized versions thereof. We will refer to the first term on the right side of (4) as the kinetic energy $`๐`$ and to the second term as the interaction energy $`๐^{\mathrm{bare}}`$.
Our choice of the product of the two normal ordered factors in the interaction, namely $`:\rho (x):`$ $`:\rho (y):`$, is taken from the book of Bjorken and Drell (Eq. (15.28) without magnetic field). It is quite natural, we agree, to start with the closest possible analog to the classical energy of a field. Of course, this term is only partly normal ordered in the usual sense and therefore it does not have zero expectation value in the unperturbed vacuum or one-particle states. It is positive, however, as a Coulomb potential ought to be.
The fully normal ordered interaction (which is not positive) is defined to be
(5)
$$๐_\alpha ^{\mathrm{ren}}=\frac{\alpha }{2}d^3xd^3yW(๐ฑ,๐ฒ):\mathrm{\Psi }^{}(x)\mathrm{\Psi }^{}(y)\mathrm{\Psi }(y)\mathrm{\Psi }(x):.$$
Bear in mind that this definition entails a determination of the splitting (2) of $``$.
We introduce a normal ordered (โdressedโ) Hamiltonian
(6)
$$_\alpha ^{\mathrm{ren}}:=d^3x:\mathrm{\Psi }^{}(x)D_{Z,m}\mathrm{\Psi }(x):+๐^{\mathrm{ren}_\alpha }.$$
(Again, this depends on the splitting of $``$.) The operator $`D_{Z,m}`$ differs from $`D_0`$ in two respects:
(7)
$$D_{Z,m}=Z^1(๐ถ๐ฉ+m\beta ),$$
where $`Z<1`$ and $`m>m_0`$. We call $`m_0`$ the bare mass and $`m`$ the โphysicalโ mass โ up to further renormalizations, which we do not address in this paper. The factor $`Z`$ is called the โwave function renormalizationโ because the โwave function renormalizationโ in the standard theory is defined by the condition that the electron-positron propagator equals $`Z`$ times a non-interacting propagator, which is the inverse of a Dirac operator. To the extent that $`๐_\alpha ^{\mathrm{ren}}`$ can be neglected, the propagator with just the first term on the right side of (6) would, indeed, be $`Z`$ times a Dirac operator with mass $`m`$. One could also interpret $`Z`$ in other ways, e.g., as a renormalization of the speed of light, but such a choice would be considerably more radical. Some other interpretations are discussed in the last section.
Our goal is to show that by a suitable choice of $`Z`$, $`m`$, and the electron space $`_+`$, we have โ up to physically unimportant infinite additive constants โ asymptotic equality between $`^{\mathrm{bare}}`$ and $`_\alpha ^{\mathrm{ren}}`$ for momenta that are much smaller than the cutoff $`\mathrm{\Lambda }`$.
(8)
$$^{\mathrm{bare}}=_\alpha ^{\mathrm{ren}}\text{ (for small momenta)}$$
The additional normal ordering in (6) will introduce some quadratic terms and it is these terms to which we turn our attention and which we would like to identify as renormalization terms. Whether our picture is really significant physically, or whether it is just a mathematical convenience remains to be seen. What it does do is indicate a need for starting with a non-conventional, non-perturbative choice of the free particle states, namely the choice of $`_+`$.
## 2. Calculation of the New Quadratic Terms
The canonical anti-commutation relations are
(9)
$$\begin{array}{cc}\hfill \{a(f),a(g)\}& =\{a^{}(f),a^{}(g)\}=\{a(f),b(g)\}=\{a^{}(f),b^{}(g)\}\hfill \\ & =\{a^{}(f),b(g)\}=\{a(f),b^{}(g)\}=0,\hfill \end{array}$$
and
(10)
$$\{a(f),a^{}(g)\}=(f,P_+g),\{b^{}(f),b(g)\}=(f,P_{}g).$$
Formally, this is equivalent to
(11)
$$\begin{array}{cc}\hfill \{a(x),a(y)\}& =\{a^{}(x),a^{}(y)\}=\{a(x),b(y)\}=\{a^{}(x),b^{}(y)\}\hfill \\ & =\{a^{}(x),b(y)\}=\{a(x),b^{}(y)\}=0,\hfill \end{array}$$
and
(12)
$$\{a(x),a^{}(y)\}=P_+(x,y),\{b^{}(x),b(y)\}=P_{}(x,y),$$
where $`P_\pm (x,y)`$ is the integral kernel of the projector $`P_\pm `$.
In order to compare the bare Hamiltonian $`^{\mathrm{bare}}`$ with the renormalized one $`_\alpha ^{\mathrm{ren}}`$ we must face a small problem; the momentum cutoff $`\mathrm{\Lambda }`$, which is necessary in order to get finite results, will spoil the theory at momenta comparable to this cutoff. Thus, our identification of the difference as a wave function and mass renormalization will be exact only for momenta that are small compared to $`\mathrm{\Lambda }`$. We will also require (although it is not clear if this is truly necessary or whether it is an artifact of our method) that $`\alpha \mathrm{ln}(\mathrm{\Lambda }/m_0)`$ is not too large.
The potential energy $`๐^{\mathrm{bare}}`$ clearly has positive singularities in it if the unregularized Coulomb interaction $`W(x,y)=|๐ฑ๐ฒ|^1\delta _{\sigma ,\tau }`$ is taken. We therefore cut off the field operators for high momenta above $`\mathrm{\Lambda }`$. (Another choice would be to cut off the Fourier transform of $`|๐ฑ๐ฒ|^1`$ at $`\mathrm{\Lambda }`$ and the result would be essentially the same; our choice is motivated partly by computational convenience.) Our cutoff procedure is to require that
(13)
$$\mathrm{\Psi }(x):=(2\pi )^{3/2}_{\{๐ฉ^3||๐ฉ|<\mathrm{\Lambda }\}}\widehat{\mathrm{\Psi }}(๐ฉ,\sigma )e^{i๐ฉ๐ฑ}๐๐ฉ$$
holds, with $`\mathrm{\Lambda }>0`$. We could also take a smooth cutoff without any significant change. This regularization keeps the positivity of $`๐^{\mathrm{bare}}`$. To make all terms finite we also need a volume cutoff for one of the difference terms. However, the volume singularity only occurs as an additive constant, which we drop. The renormalized Hamiltonian is free of infrared divergent energies and so does not need a volume cutoff.
Next, we calculate the difference
(14)
$$\begin{array}{c}:=๐^{\mathrm{bare}}๐_\alpha ^{\mathrm{ren}}=\frac{\alpha }{2}d^3xd^3yW(๐ฑ,๐ฒ)[a^{}(x)a(y)P_+(x,y)+a^{}(x)b^{}(y)P_+(x,y)\hfill \\ \hfill +b(x)a(y)P_+(x,y)b^{}(y)b(x)P_+(x,y)\\ \hfill a^{}(y)a(x)P_{}(x,y)+P_{}(x,y)P_+(x,y)+a(x)b(y)P_{}(x,y)\\ \hfill +b^{}(x)a^{}(y)P_{}(x,y)+b(x)^{}b(y)P_{}(x,y)].\end{array}$$
Thus, (4) can be rewritten as
(15)
$$\begin{array}{c}^{\mathrm{bare}}=d^3x:\mathrm{\Psi }^{}(x)D_0\mathrm{\Psi }(x):\hfill \\ \hfill +\frac{\alpha }{2}d^3xd^3y\frac{P_+(x,y)P_{}(x,y)}{|๐ฑ๐ฒ|}:\mathrm{\Psi }^{}(x)\mathrm{\Psi }(y):+๐_\alpha ^{\mathrm{ren}}\\ \hfill +\frac{\alpha }{2}d^3xd^3y\frac{P_+(x,y)P_{}(x,y)}{|๐ฑ๐ฒ|}.\end{array}$$
The first two terms on the right side are one-particle operators (i.e., quadratic in the field operators). Let us call their sum $`๐ธ`$, which we write (formally, since $`A`$ is a differential operator) as
(16)
$$๐ธ=d^3xd^3yA(x,y):\mathrm{\Psi }^{}(x)\mathrm{\Psi }(y):.$$
The last term is a cutoff dependent constant, which happens to be infinite. (One can show that this term is positive and, if we put in a momentum cutoff $`\mathrm{\Lambda }`$ and volume (infrared) cutoff $`V`$, this integral is proportional to $`V\mathrm{\Lambda }^4`$.)
We can make $`๐ธ`$ positive by choosing the normal ordering appropriately. That is, we choose $`P_+`$ to be the projector onto the positive spectral subspace of the operator $`A`$. Thus, we are led to the nonlinear equation for the unknown $`A`$
(17)
$$A=D_0+\frac{\alpha }{2}R,$$
where the operator $`R`$ has the integral kernel
(18)
$$R(x,y):=\frac{P_+(x,y)P_{}(x,y)}{|๐ฑ๐ฒ|}=\frac{(\mathrm{sgn}A)(x,y)}{|๐ฑ๐ฒ|},$$
with $`\mathrm{sgn}t`$ being the sign of $`t`$.
For physical reasons and to simplify matters we restrict the search for a solution to (17) to translationally invariant operators, i.e., $`4\times 4`$ matrix-valued Fourier multipliers. Moreover, we make the ansatz
(19)
$$A(๐ฉ):=๐ถ\omega _๐ฉg_1(|๐ฉ|)+\beta g_0(|๐ฉ|)$$
with real functions $`g_1`$ and $`g_0`$, where $`\omega _๐ฉ`$ is the unit vector in the $`๐ฉ`$ direction. In other words, we try to make $`A`$ look as much as possible like a Dirac operator. With this ansatz we have $`\mathrm{sgn}A(๐ฉ)=A(๐ฉ)/(g_1(|๐ฉ|)^2+g_0(|๐ฉ|)^2)^{1/2}`$. Thus, recalling (13), (17) can be fulfilled, if
(20) $`g_0(|๐ฉ|)`$ $`:=m_0+{\displaystyle \frac{\alpha }{4\pi ^2}}{\displaystyle _{|๐ช|<\mathrm{\Lambda }}}๐๐ช{\displaystyle \frac{1}{|๐ฉ๐ช|^2}}{\displaystyle \frac{g_0(|๐ช|)}{(g_1(|๐ช|)^2+g_0(|๐ช|)^2)^{1/2}}}`$
(21) $`g_1(|๐ฉ|)`$ $`:=|๐ฉ|+{\displaystyle \frac{\alpha }{4\pi ^2}}{\displaystyle _{|๐ช|<\mathrm{\Lambda }}}๐๐ช{\displaystyle \frac{\omega _๐ฉ\omega _๐ช}{|๐ฉ๐ช|^2}}{\displaystyle \frac{g_1(|๐ช|)}{(g_1(|๐ช|)^2+g_0(|๐ช|)^2)^{1/2}}}`$
is solvable. Note that the bare mass $`m_0`$ appears in (20).
## 3. Determination of the Dressed Electron
We will find a solution of the system (20), (21) by a fixed point argument. To this end we first integrate out the angle on the right hand side. Setting $`u:=|๐ฉ|`$ and $`v:=|๐ช|`$ we get
(22) $`g_0(u)`$ $`=m_0+{\displaystyle \frac{\alpha }{2\pi }}{\displaystyle _0^\mathrm{\Lambda }}๐v{\displaystyle \frac{v}{u}}Q_0\left({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{u}{v}}+{\displaystyle \frac{v}{u}}\right)\right){\displaystyle \frac{g_0(v)}{(g_1(v)^2+g_0(v)^2)^{1/2}}}`$
(23) $`g_1(u)`$ $`=u+{\displaystyle \frac{\alpha }{2\pi }}{\displaystyle _0^\mathrm{\Lambda }}๐v{\displaystyle \frac{v}{u}}Q_1\left({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{u}{v}}+{\displaystyle \frac{v}{u}}\right)\right){\displaystyle \frac{g_1(v)}{(g_1(v)^2+g_0(v)^2)^{1/2}}}`$
where, for $`z>1`$,
(24)
$$Q_0(z)=\frac{1}{2}\mathrm{log}\frac{z+1}{z1}$$
and
(25)
$$Q_1(z)=\frac{z}{2}\mathrm{log}\frac{z+1}{z1}1.$$
The case of zero bare mass is particularly easy. We can choose $`g_0=0`$, in which case $`g_1`$ is obtained by integration. We get
$`g_1(u)=`$ $`u+{\displaystyle \frac{\alpha }{2\pi }}{\displaystyle _0^\mathrm{\Lambda }}๐v{\displaystyle \frac{v}{u}}Q_1\left({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{u}{v}}+{\displaystyle \frac{v}{u}}\right)\right)`$
(26) $`=`$ $`u+{\displaystyle \frac{\alpha }{2\pi }}u{\displaystyle _0^{\mathrm{\Lambda }/u}}๐vvQ_1\left(\left(1/v+v\right)/2\right)`$
$`=`$ $`u+{\displaystyle \frac{\alpha }{2\pi }}u[{\displaystyle \frac{2}{3}}\mathrm{log}|(\mathrm{\Lambda }/u)^21|{\displaystyle \frac{1}{3}}(\mathrm{\Lambda }/u)^2`$
$`+{\displaystyle \frac{\mathrm{\Lambda }}{6u}}\mathrm{log}\left|{\displaystyle \frac{(\mathrm{\Lambda }/u)+1}{(\mathrm{\Lambda }/u)1}}\right|(3+(\mathrm{\Lambda }/u)^2)]`$
The function $`g_1`$ behaves asymptotically for small $`u`$ or large $`\mathrm{\Lambda }`$ as $`g_1(u)=\frac{\alpha }{2\pi }\frac{4}{3}u\mathrm{log}(\mathrm{\Lambda }/u)`$. Unfortunately, because of the $`\mathrm{log}u`$, the operator $`A`$ can never look like the renormalized Dirac operator in (7). This asymptotic expansion can be seen either from (26) or else by noting that the integrand is a continuous function, except for $`v=1`$, that decays at infinity as $`4/(3v)`$. This follows from the large $`v`$ expansion
(27)
$$vQ_1\left(\left(1/v+v\right)/2\right)=\frac{4}{3}v^1+\frac{8}{15}v^3+O(v^5).$$
For positive bare masses we solve the system (22) and (23) by a fixed point argument. To this end we define the following set of pairs of functions, for $`ฯต,\delta >0`$.
(28)
$$S_{ฯต,\delta }:=\{๐ =(g_0,g_1)|_{u[0,\mathrm{\Lambda }]}๐ (u)[m_0,(1+\delta )m_0]\times [u,(1+ฯต)u]\}.$$
Note that with the metric generated by the sup norm $`S_{ฯต,\delta }`$ is a complete metric space. Next define $`T:S_{ฯต,\delta }S_{ฯต,\delta }`$ by the right hand side of (23) and (22), i.e.,
(29) $`T_0(๐ )(u)`$ $`=m_0+{\displaystyle \frac{\alpha }{2\pi }}{\displaystyle _0^\mathrm{\Lambda }}๐v{\displaystyle \frac{v}{u}}Q_0\left({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{u}{v}}+{\displaystyle \frac{v}{u}}\right)\right){\displaystyle \frac{g_0(v)}{(g_1(v)^2+g_0(v)^2)^{1/2}}}`$
(30) $`T_1(๐ )(u)`$ $`=u+{\displaystyle \frac{\alpha }{2\pi }}{\displaystyle _0^\mathrm{\Lambda }}๐v{\displaystyle \frac{v}{u}}Q_1\left({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{u}{v}}+{\displaystyle \frac{v}{u}}\right)\right){\displaystyle \frac{g_1(v)}{(g_1(v)^2+g_0(v)^2)^{1/2}}}.`$
###### Lemma 1.
If $`Y`$, defined by
(31)
$$Y:=\alpha \mathrm{arsinh}(\mathrm{\Lambda }/m_0)/\pi ,$$
satisfies $`Y<9/50`$, if $`ฯต50Y/(950Y)`$, and if $`\delta Y/(1Y)`$, then T maps $`S_{ฯต,\delta }`$ into $`S_{ฯต,\delta }`$.
Note that $`\mathrm{arsinh}(x)=\mathrm{log}(x+\sqrt{x^2+1})`$, i.e., $`Y`$ grows logarithmically in $`\mathrm{\Lambda }/m_0`$.
###### Proof.
Obviously, $`T_0(๐ )(u)m_0`$ and $`T_1(๐ )(u)u`$.
To bound $`g_0`$ we return to (20) and note the bound (which holds in $`S_{ฯต,\delta }`$)
(32)
$$\begin{array}{cc}\hfill g_0(|๐ฉ|)& m_0+\frac{\alpha }{4\pi ^2}_{|๐ช|<\mathrm{\Lambda }}๐๐ช\frac{1}{|๐ฉ๐ช|^2}\frac{(1+\delta )m_0}{(|๐ช|^2+m_0^2)^{1/2}}\hfill \\ & m_0+\frac{\alpha }{4\pi ^2}_{|๐ช|<\mathrm{\Lambda }}๐๐ช\frac{1}{๐ช^2}\frac{(1+\delta )m_0}{(|๐ช|^2+m_0^2)^{1/2}}\hfill \\ & =m_0[1+\frac{\alpha }{\pi }(1+\delta )\mathrm{arsinh}(\mathrm{\Lambda }/m_0)].\hfill \end{array}$$
The second inequality holds because the convolution of two symmetric decreasing functions is symmetric decreasing.
Next we turn to $`g_1`$, which is a bit more complicated. We split the integration region in (21) into $`A=\{๐ช||๐ช|<2|๐ฉ|\mathrm{and}|๐ช|<\mathrm{\Lambda }\}`$ and $`B=\{๐ช||๐ช|2|๐ฉ|\mathrm{and}|๐ช|<\mathrm{\Lambda }\}`$. In region $`A`$ we use $`\omega _๐ฉ\omega _๐ช1`$ and hence the contribution to $`g_1`$ from this region is bounded above by
(33)
$$\frac{\alpha }{4\pi ^2}_{|๐ช|<\mathrm{\Lambda }}d๐ช\frac{1}{|๐ฉ๐ช|^2}\frac{(1+ฯต)2|๐ฉ|}{(|๐ช|^2+m_0^2)^{1/2}}2|๐ฉ|\frac{\alpha }{\pi }(1+ฯต)\mathrm{arsinh}(\mathrm{\Lambda }/m_0)],$$
for the same reason as in (32).
In region $`B`$ we use that $`(|๐ฉ|^2|๐ช|^2)^29|๐ช|^4/16`$. We also note, for the integration over $`B`$, that we can take the mean of the integrand for $`๐ช`$ and $`๐ช`$. In other words, we can bound this term as follows:
(34)
$$\begin{array}{cc}& \frac{\alpha }{4\pi ^2}\left|_B๐๐ช\frac{\omega _๐ฉ\omega _๐ช}{|๐ฉ๐ช|^2}\frac{g_1(|๐ช|)}{(g_1(|๐ช|)^2+g_0(|๐ช|)^2)^{1/2}}\right|\hfill \\ \hfill & \frac{\alpha }{8\pi ^2}\left|_B๐๐ช\left(\frac{\omega _๐ฉ\omega _๐ช}{|๐ฉ๐ช|^2}\frac{\omega _๐ฉ\omega _๐ช}{|๐ฉ+๐ช|^2}\right)\frac{g_1(|๐ช|)}{(g_1(|๐ช|)^2+g_0(|๐ช|)^2)^{1/2}}\right|\hfill \\ \hfill & \frac{\alpha }{8\pi ^2}_B๐๐ช\frac{4|๐ฉ||๐ช|}{(|๐ฉ|^2|๐ช|^2)^2}\frac{(1+ฯต)|๐ช|}{(|๐ช|^2+m_0^2)^{1/2}}\hfill \\ \hfill & \frac{8\alpha (1+ฯต)}{9\pi ^2}|๐ฉ|_B๐๐ช\frac{1}{|๐ช|^2}\frac{1}{(|๐ช|^2+m_0^2)^{1/2}}\hfill \\ \hfill & \frac{32\alpha (1+ฯต)}{9\pi }|๐ฉ|\mathrm{arsinh}(\mathrm{\Lambda }/m_0).\hfill \end{array}$$
Thus, we obtain the bound
(35)
$$g_1(|๐ฉ|)|๐ฉ|\{1+\frac{50\alpha }{9\pi }(1+ฯต)\mathrm{arsinh}(\mathrm{\Lambda }/m_0)\}.$$
###### Theorem 1.
If $`Y:=\alpha \mathrm{arsinh}(\mathrm{\Lambda }/m_0)/\pi <9/50`$, if $`(2+ฯต+\delta )Y<1`$ and if $`ฯต,\delta `$ satisfy the conditions of Lemma 1 then $`T`$ is a contraction.
###### Proof.
Thanks to Lemma 1 we only need to establish the contraction property. We first note that for positive real numbers $`x,y,\stackrel{~}{x},\stackrel{~}{y}`$
(36)
$$\left|\frac{x}{(x^2+y^2)^{1/2}}\frac{\stackrel{~}{x}}{(\stackrel{~}{x}^2+\stackrel{~}{y}^2)^{1/2}}\right|\frac{|\eta |}{\xi ^2+\eta ^2}((x\stackrel{~}{x})^2+(y\stackrel{~}{y})^2)^{1/2}$$
where $`(\xi ,\eta )`$ is some point on the line between $`(x,y)`$ and $`(\stackrel{~}{x},\stackrel{~}{y})`$.
Thus we get
(37)
$$\begin{array}{cc}& |[T_0(g)T_0(\stackrel{~}{g})](u)|+|[T_1(g)T_1(\stackrel{~}{g})](u)|\hfill \\ \hfill & \frac{\alpha }{2\pi }_0^\mathrm{\Lambda }dv\frac{v}{u}\left[Q_0\left(\frac{1}{2}(\frac{u}{v}+\frac{v}{u})\right)\right|\frac{g_0(v)}{\sqrt{g_0(v)^2+g_1(v)^2}}\frac{\stackrel{~}{g}_0(v)}{\sqrt{\stackrel{~}{g}_0(v)^2+\stackrel{~}{g}_1(v)^2}}|\hfill \\ & +Q_1\left(\frac{1}{2}(\frac{u}{v}+\frac{v}{u})\right)|\frac{g_1(v)}{\sqrt{g_0(v)^2+g_1(v)^2}}\frac{\stackrel{~}{g}_1(v)}{\sqrt{\stackrel{~}{g}_0(v)^2+\stackrel{~}{g}_1(v)^2}}|]\hfill \\ \hfill & \frac{\alpha }{2\pi }_0^\mathrm{\Lambda }๐v\frac{v}{u}\left[Q_0\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{(1+ฯต)v}{v^2+m_0^2}+Q_1\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{(1+\delta )m_0}{v^2+m_0^2}\right]\hfill \\ & \times \left[(g_0(v)\stackrel{~}{g}_0(v))^2+(g_1(v)\stackrel{~}{g}_1(v))^2\right]^{1/2}\hfill \\ \hfill & \frac{\alpha }{2\pi }_0^\mathrm{\Lambda }dv\frac{v}{u}[Q_0\left(\frac{1}{2}(\frac{u}{v}+\frac{v}{u})\right)\frac{(1+ฯต)v}{v^2+m_0^2}\hfill \\ & +Q_1\left(\frac{1}{2}(\frac{u}{v}+\frac{v}{u})\right)\frac{(1+\delta )m_0}{v^2+m_0^2}]๐ \stackrel{~}{๐ }\hfill \\ \hfill & \frac{\alpha }{2\pi }_0^\mathrm{\Lambda }๐v\frac{v}{u}Q_0\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{(1+ฯต)v+(1+\delta )m_0}{v^2+m_0^2}๐ \stackrel{~}{๐ }\hfill \\ \hfill & (2+ฯต+\delta )Y๐ \stackrel{~}{๐ }\hfill \end{array}$$
where $`Y=\alpha \mathrm{arsinh}(\mathrm{\Lambda }/m_0)/\pi `$ as in Lemma 1. In the last line we have simply noted that the integrals are smaller than the corresponding integral in (32). โ
###### Corollary 1.
The map $`T`$ has a unique fixed point, if $`Y<1/7`$.
###### Proof.
Take $`ฯต=50Y/(950Y)`$ and $`\delta =Y/(1Y)`$ so that $`T`$ maps $`S_{ฯต,\delta }`$ into itself. Then, if $`Y<Y_0`$, the contraction condition will be satisfied. โ
For $`\alpha =1/137`$, i.e., the physical value of the fine structure constant โ the condition $`Y<1/7`$ is fulfilled for $`\mathrm{\Lambda }/me^{18}`$.
## 4. Properties of the Renormalized Hamiltonian
What we have shown up to now is that the bare Hamiltonian in (4) is equivalent, apart from some additive constants, to a renormalized Hamiltonian. This renormalized Hamiltonian has the form (6), but the quadratic term is only approximately the one given in (7). What takes the place of $`D_{Z,m}`$ is given in (17) and (19). We see immediately from (22), (23) that the solution is a continuous pair of functions. We also see that as long as $`m_0`$ is not zero the functions $`g_0`$ and $`g_1`$ are positive and behave properly for small $`|๐ฉ|`$, i.e., $`g_0`$ is constant and $`g_1`$ is proportional to $`|๐ฉ|`$. To relate this to $`D_{Z,m}`$ we first factor out the constant $`lim_{u0+}(g_1(u)/u)`$ and call this $`Z^1`$. As is evident from (21), this constant $`1/Z`$ is larger than one โ as it should be.
The next thing to verify is that $`Zg_0(0)`$ is bigger than $`m_0`$, since this is the renormalized mass $`m`$ that appears in (7). In other words, we have to verify that
(38)
$$\frac{m}{m_0}Z^1=\frac{g_0(0)}{m_0}>\underset{๐ฉ0}{lim}\frac{g_1(|๐ฉ|)}{|๐ฉ|}=Z^1.$$
We shall, in fact, prove more than this:
###### Theorem 2.
Assuming that $`Y<1/7`$, the unique solution to the equation (20) and (21) mentioned has the property
(39)
$$\frac{g_0(|๐ฉ|)}{m_0}>\frac{g_1(|๐ฉ|)}{|๐ฉ|}$$
for all $`๐ฉ\mathrm{๐}`$.
###### Proof.
Define $`\stackrel{~}{S}_{\delta ,ฯต}`$ to be the subset of $`S_{\delta ,ฯต}`$ on which
(40)
$$\frac{g_0(|๐ฉ|)}{m_0}\frac{g_1(|๐ฉ|)}{|๐ฉ|}$$
for all $`๐ฉ\mathrm{๐}`$. If we can show that $`T`$ also leaves $`\stackrel{~}{S}_{\delta ,ฯต}`$ invariant, then we have shown the wanted inequality (39) on the solution in $`S_{\delta ,ฯต}`$ (by the uniqueness of the solution on $`S_{\delta ,ฯต}`$) except for the possibility that equality also can occur. This is so, since we can apply the fixed point argument not only to $`S_{\delta ,ฯต}`$ but also to $`\stackrel{~}{S}_{\delta ,ฯต}`$.
Showing that (40) holds is โ according to (22) and (23) โ equivalent to showing that
(41)
$$\begin{array}{c}_0^\mathrm{\Lambda }๐v\frac{v}{u}Q_0\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{g_0(v)/m_0}{(g_1(v)^2+g_0(v)^2)^{1/2}}\hfill \\ \hfill >_0^\mathrm{\Lambda }๐v\frac{v}{u}Q_1\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{g_1(v)/u}{(g_1(v)^2+g_0(v)^2)^{1/2}}\end{array}$$
holds. Now, using the Inequality (40) and the fact that the factors with the roots are monotone functions $`g_0`$ and $`g_1`$ respectively, it is enough to show that
(42)
$$\begin{array}{c}_0^\mathrm{\Lambda }๐v\frac{v}{u}Q_0\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{1}{(m_0^2+v^2)^{1/2}}\hfill \\ \hfill >_0^\mathrm{\Lambda }๐v\frac{v}{u}Q_1\left(\frac{1}{2}\left(\frac{u}{v}+\frac{v}{u}\right)\right)\frac{v/u}{(m_0^2+v^2)^{1/2}}\end{array}$$
holds. To proceed, we now compare the integrands pointwise. Using the explicit expressions (24) and (25) for $`Q_0`$ and $`Q_1`$, and with $`t=v/u`$, we will have shown that the integrand of the left hand side pointwise majorizes the one on the right hand side of (42) if
(43)
$$\mathrm{log}\left|\frac{t+1}{t1}\right|<\frac{2t}{|t^21|}$$
holds for all $`t1`$. By symmetry, we only have to consider (43) for $`t>1`$. To this end we exponentiate (43)
$$\frac{t+1}{t1}<\mathrm{exp}\left(\frac{2t}{t^21}\right)$$
and expand the exponential up to second order. (Note that this gives a lower bound on the exponential because the argument of the exponential function is positive.) I.e., it suffices to show that
(44)
$$\frac{t+1}{t1}<1+\frac{2t}{t^21}+2\frac{t^2}{(t^21)^2}$$
which follows by direct computation.
Having established (2) we see that (44) indeed gives strict inequality in (2) for the unique fixed point in $`\stackrel{~}{S}_{\delta ,ฯต}`$. โ
## 5. Interpretations of our Results
What we have done is start with the โbareโ Hamiltonian $`^{\mathrm{bare}}`$ in (4), in which the interaction is the closest analog to the classical electrostatic energy of a field. After some analysis, we found in section 3 that $`Z`$ and $`m`$ could be uniquely chosen so that for momenta much less than the ultraviolet cutoff $`\mathrm{\Lambda }`$, $`^{\mathrm{bare}}=_\alpha ^{\mathrm{ren}}`$ plus a well defined infinite constant. Of course the field $`\mathrm{\Psi }`$ in the two cases is different. Formally, they differ by a Bogolubov transformation, but in fact an inequivalent, $`\alpha `$ dependent, representation of the CAR is needed. In section 4 we showed that not only is $`Z^1m>m_0`$ but more is true, namely $`m>m_0`$, and this is comforting physically.
We emphasize that we have not โintegrated outโ any field variables. Our new Hamiltonian $`_\alpha ^{\mathrm{ren}}`$ is the same as the original one โ on a purely formal level. It is suggestive, nevertheless, that a good deal of the electrostatic energy has now been incorporated into the leading term $`d^3x:\mathrm{\Psi }^{}(x)D_{Z,m}\mathrm{\Psi }(x):`$ and that the remaining interaction is somehow less important than the original one. It is, after all, normal ordered, which means it vanishes on one-electron states, if we define such states by $`\mathrm{\Psi }^{}|0`$, where $`|0`$ is the vacuum of the new $`\mathrm{\Psi }`$. While this makes sense perturbatively, it is, however, misleading because the new vacuum (the state of lowest energy) is surely not the obvious choice $`|0`$.
If we drop the new interaction term $`๐_\alpha ^{\mathrm{ren}}`$ we are left with $`d^3x:\mathrm{\Psi }^{}(x)D_{Z,m}\mathrm{\Psi }(x):`$ as our Hamiltonian. Unfortunately this is not the Hamiltonian of a Dirac operator (even at low momenta) because of the factor $`Z^1`$. We have called this a wave function renormalization, but that is not really in the spirit of renormalizing the one-electron states (which is what is usually done) and is, instead, a renormalization of an operator. One school, (Kรคllรฉn ) thinks it is proper to speak of a renormalized $`\mathrm{\Psi }`$ by incorporating a factor $`Z^{1/2}`$ into $`\mathrm{\Psi }`$, but this changes the anti-commutation relations! Note that this formulation requires renormalizing the bare mass $`m_0`$ to $`g_0(\mathrm{๐})/Z`$ and renormalizing the bare charge $`e`$ to $`e/Z`$.
Another point of view is to regard $`Z^1m`$ as the physical mass and to change $`Z^1๐ฉ`$ into $`๐ฉ`$ by a unitary transformation (which is nothing other than a change of length scale). This has the disadvantage of changing the speed of light or Planckโs constant. It also would mean a different scale change for particles of different mass, e.g., muons. As opposed to Kรคllรฉnโs procedure, this requires renormalizing the mass from $`m_0`$ to $`g_0(\mathrm{๐})`$ only, but there is no need for charge renormalization.
Another possibility is to bring out the factor $`Z^1`$ as a multiplier of the whole Hamiltonian, which would mean changing the fine structure constant to $`\alpha Z`$, i.e., a charge renormalization but now from $`e`$ to $`e/Z^{1/2}`$ only. The obvious problem here is that since the Hamiltonian is the generator of time translation, this means a change of the time scale (which, again, depends on the particle in question).
Doubtless, different people will have different opinions about these matters. We do not wish to commit ourselves to any point of view. But it is our opinion that the construction of $`_\alpha ^{\mathrm{ren}}`$ is a significant piece of the puzzle of constructing a nonperturbative QED.
Acknowledgement: The authors thank Dirk Hundertmark for a useful discussion. Financial support of the European Union, TMR grant FMRX-CT 96-0001, the U.S. National Science Foundation, grant PHY98-20650, and NATO, grant CRG96011 is acknowledged. |
warning/0003/astro-ph0003031.html | ar5iv | text | # ISO Spectroscopy of the HH 7-11 Flow and its red-shifted counterpart0footnote 00footnote 0
## 1. Introduction
Atomic and molecular outflows trace the mass loss from protostellar objects, which is a fundamental characteristic of the formation and evolution of low mass stars. These outflows are often traced by the optical Herbig-Haro (HH) objects, shock-excited nebulosities which mark the interface between outflowing and circumstellar material. One of these systems, which since its discovery (Herbig H74 (1974) ; Strom et al. Setal74 (1974)) has been subjected to a detailed multi-wavelength analysis, is HH 7-11. The system is relatively bright and lies in the very active star forming NGC 1333 region (Aspin et al. ASR94 (1994); Bally et al. Betal96 (1996)). A distance to the outflow of 350pc (Herbig & Jones HJ83 (1983)) is widely adopted in the literature, although Cernis (C93 (1993)) proposes 200pc. Early optical spectroscopic studies show that the HH 7-11 outflow has a complex velocity field and low excitation (Solf & Bรถhm SB87 (1987); Bรถhm & Solf BS90 (1990)), a conclusion further supported by near infrared studies (Hartigan et al. HCR89 (1989); Carr Carr93 (1993)). Early on, a well defined CO bipolar outflow was detected associated with this system (Snell & Edwards SE81 (1981); Bachiller & Cernicharo BC90 (1990)), which also is observed in some other molecules like HCO<sup>+</sup> and H<sub>2</sub>O (Mehringer M96 (1996); Cernicharo et al. CBG96 (1996)).
The HH 7-11 outflow has an unusual morphology; the blue outflow lobe is made up of an arc-shaped chain of knots, while the red one is invisible at optical wavelengths. The red lobe is, however, detected in the (1-0)S(1) H<sub>2</sub> line at 2.12$`\mu `$m having a very ragged appearance (see Fig 1). Both lobes have a total extension of $``$ 2. It has been thought that the driving source of the outflow is the infrared star SVS 13 (Strom et al. Setal76 (1976)), a conclusion partially supported by the proper motion measurements of the knots (Herbig & JonesHJ83 (1983)) and the source observed outbursts (Goodrich G86 (1986); Eislรถffel et al. Eetal91 (1991); Liseau et al. LLM92 (1992)). The source has a luminosity of $`85`$ $`L_{}`$ (Molinari et al. MLL93 (1993)). Recent high angular resolution ($``$ 0.<sup>โฒโฒ</sup>3) VLA continuum observations at 3.6 cm suggest another nearby embedded source (VLA 3) as a likely candidate, based on its better alignment with the HH string. Interferometric observations (Bachiller et al. Betal98 (1998)) at 1.3 and 3.5 mm with better than $`0.\mathrm{}2`$ resolution, however, have not confirmed this. The interferometric observations have led to the discovery of a second jet emanating from a more deeply embedded source 14.<sup>โฒโฒ</sup>5 away from SVS 13, named SVS 13B (Grossman et al. Getal87 (1987)).
One of the reasons why the HH 7-11 flow has been so intensively studied (see e.g. Reipurth R94 (1994)), is that it was the first system showing clear signatures of a high velocity outflow in both neutral and molecular gas tracers (Lizano et al. Letal88 (1988); Rodriguez et al. Retal90 (1990); Giovanardi et al. Getal92 (1992)). This was a major step forward in the interpretation of molecular outflows as being driven by faster but more tenuous (than the outflow entrained gas) atomic stellar winds (Masson & Chernin MC93 (1993); Raga et al. Retal93 (1993)). This has led to a more careful analysis of the energetics and shock conditions associated with the ionic/atomic and molecular gas outflows (Raga Ra91 (1991)). HH 7-11 is also one of the few examples where it is possible to disentangle the contributions of shock excited and fluorescent emission from its near infrared H<sub>2</sub> spectra (Gredel G96 (1996); Fernandes & Brand FB95 (1995); Everett E97 (1997)).
In the present study, we take advantage of the capabilities of the Infrared Satellite Observatory (ISO, Kessler et al. Ketal96 (1996)) spectrometers to study the mid- and far infrared emission line spectra from the HH 7-11 red and blue lobes and around the driving source SVS 13. The observations are described in Sect. 2; the results are presented and discussed in Sect. 3 and following, and the main conclusions are summarised in Sect. 9.
## 2. Observations and Data Reduction
We used the two spectrometers on the ISO satellite to observe several locations along the HH 7-11 flow, its optically invisible counterflow and the candidate exciting source SVS 13. The Long Wavelength Spectrometer (LWS, Clegg et al. Clegg96 (1996)) was used in its LWS01 grating mode to acquire full low resolution (R $``$ 200) 43-197$`\mu `$m scans with data collected every 1/4 of a resolution element (equivalent to $``$0.07$`\mu `$m for $`\lambda <`$90$`\mu `$m, and to $``$0.15$`\mu `$m for $`\lambda >`$90$`\mu `$m); a total of 19 scans were collected, corresponding to 38s integration time per spectral element. The LWS was also used in LWS04 Fabry-Perot (FP) mode to collect high (R $``$ 8000) resolution scans of the \[Oi\]63$`\mu `$m line; 52 scans sampled at 1/4 of the resolution element (i.e., $``$ 0.0017 $`\mu `$m, or 8 km s<sup>-1</sup>) were collected, equivalent to an integration time of $``$ 100s per spectral element. The Short Wavelength Spectrometer (SWS, de Graauw et al. DGetal96 (1996)) was used in its SWS02 grating mode to observe line scans at medium resolution (R $``$ 2000) for selected wavelength regions covering the \[Siii\]34.8$`\mu `$m and \[Neii\]12.8$`\mu `$m fine structure lines, and the pure rotational transitions of molecular hydrogen from (0-0)S(1) to (0-0)S(7). All relevant information is summarised in Table 1, including the Astronomical Observation Templates used, the revolution number and the total observing time of the observations.
LWS data processed through Off-Line Processing (OLP), version 7, have been reduced using the LWS Interactive Analysis<sup>1</sup><sup>1</sup>1LIA is available at http://www.ipac.caltech.edu/iso/lws/lia/lia.html (LIA) Version 7.2. The dark current and gain for each detector were re-estimated, and the data were recalibrated in wavelength, bandpass and flux. The absolute flux calibration quoted for LWS in grating mode is 10-15% and it is valid for point-like sources since the primary calibrator, Uranus, is point-like to the LWS beam; however, our sources are not rigorously point-like and we adopt a more conservative number of 20%. Additional processing for the FP data (with the LIA routine FP\_PROC) included a gain correction to compensate for the incorrect positioning of the grating during the FP observations. The integrated \[Oi\]63$`\mu `$m line fluxes obtained with the FP are about a factor of 2 lower than the values measured using the grating. This discrepancy is larger than the 30% figure generally quoted for the absolute flux calibration of the FP (Swinyard et al Swin98 (1998)). We cannot offer any explanation for this difference, and in this paper we will not use the line fluxes measured with the FP. The accuracy of the FP wavelength calibration is believed to be better than 1/2 resolution element, or $``$ 15 km s<sup>-1</sup>.
SWS data were processed using OSIA, the SWS Interactive Analysis<sup>2</sup><sup>2</sup>2OSIA is available at http://www.mpe.mpg.de/www\_ir/ISO/observer/osia/osia.html . Dark currents and photometric checks were revised; in many cases the former were corrupted on a few detectors and were re-estimated and subtracted. The March 1998 bandpass calibration files have been used to produce the final spectra. The absolute flux calibration for SWS data should be accurate to within 20%.
The final steps of data analysis were done using the ISO Spectral Analysis Package<sup>3</sup><sup>3</sup>3ISAP is available at http://www.ipac.caltech.edu/iso/isap/isap.html (ISAP) Version 1.5 for both LWS and SWS. Grating scans (LWS) and detectors spectra (SWS) were averaged using a median clipping algorithm optimised to flag and discard outliers mainly due to transients; line fluxes were estimated by means of gaussian fitting (multiple gaussians in case of blended lines). The LWS observations toward HH 7 and the red lobe (RL 1+2) were heavily fringed due to the vicinity of the relatively strong continuum source SVS 13; standard techniques available under ISAP were used to remove these instrumental effects.
The locations of the observed positions and instrument apertures are shown in Fig. 1 superimposed on an H<sub>2</sub> (1-0)S(1) 2.12$`\mu `$m continuum-subtracted map obtained with the near infrared camera at the 60<sup>โฒโฒ</sup> Mt. Palomar telescope (Murphy et al. Metal95 (1995)). The apertures of the two ISO spectrometers are quite different. The SWS focal plane aperture is 14<sup>โฒโฒ</sup>x20<sup>โฒโฒ</sup> for all of the detected lines except for \[Siii\]34.8$`\mu `$m (20<sup>โฒโฒ</sup>x33<sup>โฒโฒ</sup>) so that, at each pointing, the contamination from nearby knots should be negligible. The LWS aperture is rather large ($`ร`$80<sup>โฒโฒ</sup>, Swinyard et al. Swin98 (1998)); the pointing on SVS 13 also includes HH 11 and 10 and two sources VLA 3 and SVS 13B (see Sect. 1), while that centered on HH 7 also includes HH 8, 9 and 10.
## 3. Results
Table 2 presents the fluxes for the lines detected toward each location. One sigma upper limits are given only for lines observed with dedicated SWS02 โline scansโ AOTs but not detected (see Tab. 1); a horizontal dash means that no observation is available for that particular line.
Detected lines are also plotted in Figs. 2a, b, c and d; most of them are detected everywhere. Exceptions are \[Siii\]34.8$`\mu `$m, which is not detected toward RL 2, ortho-H<sub>2</sub>O which is not detected toward RL 1+2, and \[Neii\]12.8$`\mu `$m which is not detected anywhere. All lines, except \[Siii\]34.8$`\mu `$m and the H<sub>2</sub> lines, are stronger on the SVS 13 position.
The FP spectra of the \[Oi\]63$`\mu `$m line toward HH 7 and RL 1+2 are presented in Fig. 3. The lines are resolved ($`\mathrm{\Delta }v_{FP}`$30 km s<sup>-1</sup>) with deconvolved FWHM of $``$50 km s<sup>-1</sup>and $``$25 km s<sup>-1</sup> for the two positions; these widths are consistent with the velocity field traced by the \[S II\]6717ร
line (Stapelfeldt S91 (1991)) toward HH 7, and suggest that \[Oi\]63$`\mu `$m originates from the flow material.
## 4. The Gas Physical Parameters
### 4.1. H<sub>2</sub>
The observed H<sub>2</sub> lines arise from quadrupole rotational transitions of the ground vibrational state so that they are likely to be optically thin. The radiative lifetimes range from 0.54 yrs for level $`v`$=0, J=7 where line (0-0)S(5) originates, to $`1000`$ yrs for level $`v`$=0, J=2 which is the upper state for (0-0)S(0) transition; thus these levels are collisionally populated, and a simple LTE analysis is adequate to interpret the data (e.g., Gredel G94 (1994)). Einstein coefficients and wavenumbers were taken from Black & Dalgarno (BD76 (1976)) and Dabrowski (D84 (1984)); ortho/para=3 has been assumed. We dereddened the H<sub>2</sub> line fluxes using the visual extinction given by Gredel (G96 (1996)) for the individual HH knots, and the Rieke & Lebofsky (RL85 (1985)) extinction curve. Dereddened line fluxes were used to produce the Boltzmann plots shown in Fig. 4; a single temperature fit is in good agreement with the observed fluxes at all positions. The solid angles of line emitting regions are arbitrarily set to the equivalent SWS focal plane aperture, i.e. 14<sup>โฒโฒ</sup> x 20<sup>โฒโฒ</sup> or 6.6$`\times 10^9`$ sr (valid for all detected H<sub>2</sub> lines but S(1), where the solid angle is $``$30% higher), assuming a beam filling factor of 1.
The H<sub>2</sub> temperatures and the SWS beam averaged column densities, together with 1$`\sigma `$ uncertainties are listed in Table 3. The H<sub>2</sub> column densities should be considered as lower limits because the beam filling factor can be less than 1. The H<sub>2</sub> temperatures vary by less than 15% along the flow and appear to trace an H<sub>2</sub> component (which we call โwarmโ) which is colder than the one identified by Gredel (G96 (1996)) via higher excitation H<sub>2</sub> ro-vibrational lines (2100 K$``$ T $``$2750 K, which we will call โhotโ). The column densities for the warm H<sub>2</sub> (see Table 3, col. 3) are on average two orders of magnitude higher than those of the hot H<sub>2</sub> (Gredel G96 (1996)) and the difference cannot be explained by the smaller (a factor 2) emitting areas adopted by Gredel<sup>4</sup><sup>4</sup>4We believe that Table 3 of Gredel (G96 (1996)) incorrectly reports the H<sub>2</sub> column densities; based on the explanation in note $`c`$ of that Table, the total column densities should be a factor ten higher than reported in its column 7. Furthermore (R. Gredel, priv. comm.) the IRSPEC slit was not aligned exactly on the peaks of the individual knots, suggesting that quoted column densities should more conservatively be considered as lower limits. . This suggests that the two H<sub>2</sub> components may be actually distinct.
### 4.2. CO
CO rotational lines have been detected at all three positions observed with LWS, although the intensity and number of detected lines are highest toward SVS 13. More CO lines than those reported in Table 2 are marginally visible in our spectra toward the two flow positions, but it is impossible to reliably determine their flux because of the heavily fringed LWS spectra. Contrary to the H<sub>2</sub> lines, the CO rotational spectrum arises from dipole transitions and the simplifying assumptions of optically thin lines cannot be adopted a priori. We analysed the CO lines using the LVG model described in Ceccarelli et al. (Cetal98 (1998)). Under NLTE conditions the line ratios depend both on the gas temperature and density, as well as on the CO column density if the lines are optically thick. The observed CO lines cannot constrain the three parameters simultaneously. The distribution of CO line fluxes vs J toward SVS 13 is essentially flat, and we find a range of physical conditions which are consistent with our observations. We find, as extreme cases, a โcoldโ solution with T$``$350 K and $`n10^6`$ cm<sup>-3</sup>, and a โwarmโ solution with T$``$900 K and $`n10^5`$ cm<sup>-3</sup>; the model also predicts optically thin CO lines. Interestingly, this temperature range is centered around the value independently derived from H<sub>2</sub> lines, which suggests that the CO and H<sub>2</sub> emission come from the same region. For T=560 K (the H<sub>2</sub> temperature toward SVS 13, see Table 3) we obtain $`n4\times 10^5`$ cm<sup>-3</sup>. The CO line ratios toward HH7 and RL 1+2 imply temperatures and densities similar to those around SVS 13, although the lower absolute line fluxes would suggest column densities or beam filling factor about a factor two lower. Using the SWS beam solid angle we derive N(CO)$`2\times 10^{16}`$ cm<sup>-2</sup> toward the SVS 13 LWS pointing, giving a direct measurement of the CO abundance in this warm gas of \[CO\]/\[H\]$`1.2\times 10^4`$. Based on this number it seems that CO accounts for essentially all of the gas-phase carbon in the interstellar medium (Cardelli et al. CMJS96 (1996)). Recently Lefloch et al. (Letal98 (1998)) found evidence of CO depletion toward the core of SVS 13, likely due to condensation onto grains in the dense environments close to SVS 13. We find no evidence of such a depletion in the high-J CO line emitting region, suggesting that all of the CO locked into the grain mantles has been returned to the gas-phase.
### 4.3. H<sub>2</sub>O
As far as H<sub>2</sub>O is concerned, since only the 3(0,3)-2(1,2) 174.6$`\mu `$m and 2(1,2)-1(0,1) 179.5$`\mu `$m lines of ortho-H<sub>2</sub>O have been detected in SVS 13 and HH 7, it is clear that water plays only a secondary role in gas cooling (see Table 4) compared to some other cases of outflow exciting sources such as IC 1396N (Molinari et al. Metal98 (1998)) or L1448-mm (Nisini et al. Netal99 (1999)). This result goes against earlier suggestions based on previous $`\lambda `$=1.67 mm water line observations (Cernicharo et al. CBG96 (1996)), that the emission originates in dense (10<sup>6</sup> cm<sup>-3</sup>) shocked material with a water abundance comparable to that of CO. It should be noted, however, that Cernicharo et al. assumed a gas temperature of 50 K, while our observations of H<sub>2</sub> and CO lines clearly indicate that the temperature of the molecular material is a factor of 10 higher. The maps presented by Cernicharo et al. show that water emission is concentrated within $`10`$โณ region centered on HH 11. Adopting T = 560 K and n = 4$`\times 10^5`$ cm<sup>-3</sup>, compatible with the conditions for H<sub>2</sub> and CO, our model fit (Ceccarelli et al. Cetal98 (1998)) to ISO and millimeter water lines predicts optically thick lines and a water column density N(H<sub>2</sub>O)$`10^{15}`$ cm<sup>-2</sup>, about a factor 20 lower than that estimated by Cernicharo et al. An abundance \[H<sub>2</sub>O\]/\[H<sub>2</sub>\]$`<\mathrm{\hspace{0.33em}6}\times 10^6`$ is also derived assuming that the lines from two molecules come from the same gas component; we will prove this assumption in Sect. 6.
## 5. The Photo-Dissociation Region
In spite of the shock-excited nature of the HH objects, it is likely that a non-negligible contribution to the observed line emission actually come from an extended Photo-Dissociation Region (PDR, Tielens & Hollenbach TH85 (1985)) component associated with the NGC 1333 cloud. This is suggested by low dispersion ISO-LWS observations (Caux et al. Cetal00 (2000)) at various positions in the NGC 1333 cloud, which show a widespread \[Oi\]63$`\mu `$m and \[Cii\]158$`\mu `$m emission; this extended component seems to account for $``$20% of the \[Oi\]63$`\mu `$m, and all of the \[Cii\]158$`\mu `$m emission we see from our LWS pointings. Our FP data (Sect. 2) confirm that only a small fraction of the \[Oi\]63$`\mu `$m line may come from a quiescent PDR component.
Using the PDR Toolbox<sup>5</sup><sup>5</sup>5The โPDR Toolboxโ is available at http://dustem.astro.umd.edu and contains downloadable FIR lines diagnostic information about PDRs. The tool has been created by L. Mundy, M. Wolfire, S. Lord and M. Pound, and it is based on the new PDR models of Kaufman et al. (KWHL99 (1999))., the observed \[Cii\]158$`\mu `$m emission requires a relatively faint FUV irradiation level of G$`{}_{0}{}^{}<`$10 in units of local Galactic FUV flux (Habing H68 (1968)). Although SVS 13 might be able to provide the required FUV field, the widespread \[Cii\]158$`\mu `$m emission seen by Caux et al. (Cetal00 (2000)) clearly suggest an external irradiation source. A natural candidate is BD +30 549, the B6 star responsible for the illumination of the NGC 1333 reflection nebulosity (Harvey et al. Hetal84 (1984)); located $``$ 0.8 pc N-NE of the HH 7-11 area, it can certainly provide the needed G$`{}_{0}{}^{}<\mathrm{\hspace{0.33em}10}`$ FUV field. In this regime, the PDR surface temperature does not exceed $``$ 100 K (Kaufman et al. KWHL99 (1999), Timmermann et al. Tetal96 (1996), Kemper et al. Ketal99 (1999), Liseau et al. Letal99 (1999)) which excludes a PDR origin for the \[Siii\]34.8$`\mu `$m line (Hollenbach, Takahashi & Tielens HTT91 (1991)), the H<sub>2</sub> and the CO lines. Indeed, the temperature of the molecular material is at least 5 times higher, and the molecular emission does not appear to be extended as one would expect for a PDR origin; ISOCAM-CVF near-IR imaging spectroscopy of the HH 7-11 region extracted from the public ISO data archive<sup>6</sup><sup>6</sup>6The ISO archive is available at http://pma.iso.vilspa.esa.es shows that the emission from the same H<sub>2</sub> rotational lines observed with SWS is concentrated along the flow and peaks in correspondence of visible HH objects (Noriega-Crespo et al. Nor00 (2000)), obviously favouring a shock origin for these lines.
## 6. Shocks along the Flow
It is known that the nature of the shock excitation is dramatically influenced by the presence of a magnetic field component perpendicular to the shock velocity which prepares the up-stream medium and smoothes out the effect of the front passage. The differences in the physical conditions of shocked gas are such that two distinct classes of shocks, C(ontinuous) and J(ump) have been idealised (Draine D80 (1980)). In a J-shock the temperature reached by the shocked material depends on the square of the shock velocity and can be as high as 10<sup>5</sup> K, resulting in complete molecular dissociation. In C-shocks, the temperature rarely exceeds a few thousand degrees and molecular material can survive. These two very different physical scenarios produce distinctive signatures in terms of cooling ratios between different species, and of line ratios within the same species. We will show that our data of the HH 7-11 region depict a complex situation where the two types of shock coexist; to help the discussion below, we report in Table 4 the total cooling rates in the various species as, when applicable, derived from the models used to estimate their physical parameters.
### 6.1. J-shocks
First of all we note that the non-detection of the \[Neii\]12.8$`\mu `$m line down to a level of 10<sup>-20</sup> W cm<sup>-2</sup> confirms the low-excitation nature of the HH 7-11 chain and suggests (Hollenbach & McKee HM89 (1989), hereafter HM89) a shock velocity v$`{}_{s}{}^{}<\mathrm{\hspace{0.33em}40}50`$ km s<sup>-1</sup>, depending of the pre-shock density, in excellent agreement with our high-resolution FP \[Oi\]63$`\mu `$m observations (Sect. 2) and with estimates from optical spectroscopy (Bรถhm, Brugel & Olmsted BBO83 (1983); Solf & Bรถhm SB87 (1987)). This upper limit on the shock velocity excludes strong shocks, yet the detection of \[Siii\]34.8$`\mu `$m requires the presence of a dissociative shock component, since negligible ionization is expected from a C-shock (HM89). Such a component would also explain the observed \[Oi\]63$`\mu `$m cooling, since the latter is expected to be the main coolant in J-shocks<sup>7</sup><sup>7</sup>7Copious \[Oi\]63$`\mu `$m can also be produced in non-dissociative shocks. In this case however, the presence of H<sub>2</sub> would allow the incorporation of O into water, via the chain of endothermic reactions O \+ H<sub>2</sub> $``$ OH + H and OH + H<sub>2</sub> $``$ H<sub>2</sub>O + H. This chain has an activation energy of T$``$ 220 K, lower than the H<sub>2</sub>temperature estimated from the rotational lines (between 540 and 670 K, see Table 3); we conclude that the observed \[Oi\]63$`\mu `$m emission cannot originate in C-shocks.. In order to compare the \[Oi\]63$`\mu `$m with the \[Siii\]34.8$`\mu `$m line for HH 7 and SVS 13, we need to determine the fraction of \[Oi\]63$`\mu `$m emission due to other HH objects falling within the LWS beam.
The LWS beam centered on HH 7 also contains HH 8, 9 and 10; since the excitation conditions for the different HH objects along the flow do not show dramatic variations (Hartigan, Curiel & Raymond HCR89 (1989)), we choose to use the detected \[Siii\]34.8$`\mu `$m lines toward HH 7 and HH 10 as weights to estimate the \[Oi\]63$`\mu `$m cooling intrinsic to HH 7. We conservatively assign to HH 8 and HH 9 the same \[Siii\]34.8$`\mu `$m flux measured toward HH 7, and we also consider that the weighting of the LWS beam profile decreases the contribution of HH 8-9 and HH 10 to the \[Oi\]63$`\mu `$m measured on the HH 7 pointing by 10% and 50%, respectively. Similar estimates can be done for the SVS 13 pointing, where again HH 10 contributes at a 50% level. The HH 11โs contribution does not need to be disentangled since it also contributes to the \[Siii\]34.8$`\mu `$m line of SVS 13. No correction is required for the RL 1+2 pointing. Taking the above into account, we obtain \[Oi\]63$`\mu `$m/\[Siii\]34.8$`\mu `$m line ratios between 15 and 20, suggesting a pre-shock density $`n_010^4`$ cm<sup>-3</sup> (HM89). The models can also reproduce the absolute fluxes, provided that the emission solid angles are $``$ 5<sup>โฒโฒ</sup> in diameter with filling factor of 1; more intense Oi and Siii lines on the central position are likely due to slightly larger solid angles and/or filling factors.
As concerns molecular cooling the main contribution to H<sub>2</sub> emission in J-shocks is predicted (HM89) to come from material excited by FUV or H<sub>2</sub>-formation pumping. Indeed, Fernandes & Brand (FB95 (1995)) propose a fluorescent origin for the near-IR H<sub>2</sub> lines toward HH 7. For the (0-0)S lines however, the predicted line ratios are not reproduced by the observations. In particular the S(7) line, which is predicted to be always brighter than the S(5) line irrespectively of pre-shock density and shock velocity, is not detected at all in our SWS spectra<sup>8</sup><sup>8</sup>8The above mentioned ISOCAM-CVF observations (at lower spectral resolution than the SWS) actually detected the S(7) but with a flux about 4 times lower than the S(5) line flux, compatibly with our SWS upper limits on the S(7) line (Noriega-Crespo et al. Nor00 (2000))..
### 6.2. C-shocks
When interpreted in terms of non-dissociative shock, the pure rotational H<sub>2</sub> lines provides a sensitive probe for the shock velocity, given the four orders of magnitudes of dynamical range spanned by their line ratios. This is shown in Fig. 5, which presents the observed H<sub>2</sub> line ratios superimposed on a grid of C-shock models from Kaufman & Neufeld (KN96 (1996)).
Our observations seem to place quite stringent boundaries to the shock velocity, $`15<v_s<\mathrm{\hspace{0.33em}20}`$ km s<sup>-1</sup>. The corresponding gas temperature, $`600`$ K (Kaufman & Neufeld KN96 (1996)), is in good agreement with the values derived using simple LTE analysis (Sect. 4), as expected since the molecular gas is collisionally heated by the C-shock front. Observed absolute fluxes can also be reproduced by the model as long as the emission solid angle does not exceed few arcsecs in diameter. We note that such a low velocity shock would also produce very faint S(6) and S(7) lines. Once the shock velocity is determined, we can use the CO/H<sub>2</sub> cooling ratio to estimate the pre-shock density. Fig. 6 presents such a diagnostic diagram.
The CO line fluxes in the three LWS pointings should be corrected for contamination by the other HH objects. In similar way as for the \[Oi\]63$`\mu `$m/\[Siii\]34.8$`\mu `$m ratio, we use the H<sub>2</sub> coolings in different positions as weights to split the individual contributions to the CO cooling. The corrected cooling ratios are reported as dashed lines in Fig. 6. We see that a shock velocity $`15<v_s<\mathrm{\hspace{0.33em}20}`$ km s<sup>-1</sup> would correspond to a pre-shock densities of $`n_010^4`$ cm<sup>-3</sup> for HH 7 and RL 1+2, and $`n_010^5`$ cm<sup>-3</sup> for SVS 13.
Water is an important diagnostic for C-shock models. H<sub>2</sub>O cooling is a fast function of the shock velocity since free atomic oxygen in the post-shock gas is expected to be incorporated into water as soon as the temperature rises above $`200`$ K, roughly corresponding to $`v_s10`$ km s<sup>-1</sup> (Draine, Roberge & Dalgarno DRD83 (1983), Kaufman & Neufeld KN96 (1996)). In our case then, where the temperature is $``$600 K, cooling via H<sub>2</sub>O rotational transitions is expected to be dominant with respect to that of CO. Instead, the diagnostic diagram in Fig. 7 shows that the observed CO/H<sub>2</sub>O cooling ratio is consistent with shock velocities $`v_s<\mathrm{\hspace{0.33em}12}`$ km s<sup>-1</sup>, which cannot justify the observed H<sub>2</sub> temperatures. This suggests that most of the expected gas phase water is missing. We will propose an explanation for this result in Sect. 7.1
## 7. The Shocks and the Herbig-Haro Objects
We have shown that HH 7-11 is a complex region where different line emission mechanisms are simultaneously at work. Notwithstanding the poor spatial resolution of our data, the variety of different spectral signatures detected allow us to draw the following physical scenario of the region.
### 7.1. HH 7
Starting from HH 7, the C-shock conditions diagnosed by the molecular emission clearly require the presence of a magnetic field $`๐_0`$$``$$`๐ฏ_{s_C}`$ (Draine D80 (1980)), where $`๐ฏ_{s_C}`$ is the shock velocity of the C component. Assuming the standard scaling law (Draine D80 (1980)) for the value of $`B_0`$
$$B_0bn_0^{0.5}\mu \mathrm{G}$$
(1)
where $`b1`$ in the interstellar medium, the estimated $`n_010^4`$cm<sup>-3</sup> pre-shock density for the C-shock component on HH 7 gives $`B_0`$$`100`$$`\mu `$G. Variations of a factor 3 in each direction are nevertheless possible, since $`b`$ can vary between 0.3 and 3 in molecular clouds (HM89). For similar $`B_0`$ and $`n_0`$, Draine, Roberge & Dalgarno (DRD83 (1983)) have shown that the C$``$J transition occurs at $`v_s`$$``$50 km s<sup>-1</sup>, similar to the upper limit set on $`v_{s_J}`$ (the J-shock velocity) by our observations. This means that a significant $`๐_0`$ component transverse to $`๐ฏ_{s_J}`$ would smooth the shock front to a C type or, equivalently, that our observed J-shock component can only exist as long as $`๐ฏ_{s_J}`$$``$$`๐_0`$. When the bowshock-like morphology of HH 7 is also considered, as revealed by HST NICMOS 2.12$`\mu `$m images (A. Cotera, priv. comm.), then a very simple scenario emerges where HH 7 is immersed in a $`๐_0`$ roughly parallel to the flow axis. At the tip of the bow $`๐ฏ_s`$$``$$`๐_0`$ and J-shock conditions are present; \[Siii\]34.8$`\mu `$m and most of \[Oi\]63$`\mu `$m flux arise from this region, and favourable conditions also exist to originate the โhotโ, T$`2100`$ K, H<sub>2</sub> emission detected by Gredel (G96 (1996)). Along the sides of the bow, $`๐ฏ_s`$ becomes nearly perpendicular to $`๐_0`$, creating favourable conditions for C-type shocks. This picture also provides a plausible explanation to the problem of the missing water: the water is actually produced in the C-shocks, but rapidly condenses onto dust grains and disappears from the gas-phase. This possibility is suggested by the recent discovery with ISO-LWS (Molinari et al. Metal99 (1999)) of crystalline water ice toward HH 7. The deduced water abundance (in solid state form) is comparable to the interstellar oxygen abundance, which is what the models would predict for gas-phase water in C-shocks; furthermore, the fact that the ice is in crystalline form requires grain temperatures of the order of 100 K, only attainable in dissociative shocks (HM89; Draine, Roberge & Dalgarno DRD83 (1983)). After being heated, the grains would be efficiently transported in the post C-shock regions along the B lines, directed parallel to the flow. For the gas parameters we derived, the grains are efficiently coupled to the magnetic field and are not significantly decelerated by collisions with the neutrals (Draine, Roberge & Dalgarno DRD83 (1983)).
### 7.2. HH 10
From the viewpoint of the line emission properties, HH 10 and HH 7 appear very similar objects. Although no LWS data were specifically collected toward HH 10, the H<sub>2</sub> and the \[Siii\]34.8$`\mu `$m lines still argue in favour of a dual J+C shock nature. From the morphological point of view, HH 10 appears as an irregular blob in the 2.12$`\mu `$m image of Fig. 1. Higher spatial resolution images in H$`\alpha `$ and \[Sii\]6717+31ร
from HST (unpublished archival data) resolved HH 10 into a double filamentary structure whose N-S orientation does not appear to be related to the flow direction. However, the H$`\alpha `$/\[Sii\]6717+31ร
emission ratio is higher in the NW part of this HH object, facing toward SVS 13. A higher ratio implies higher excitation (stronger shocks) conditions, which are likely to be traced by the \[Siii\]34.8$`\mu `$m line and by the โhotโ, T$`2200`$ K, H<sub>2</sub> component (Gredel G96 (1996)). If we assume that $`๐_0`$ maintains its direction downstream along the flow, as suggested for HH 7 (see above), then we speculate that the morphology of HH 10 could correspond to irregularities or corrugations in the shocked walls of the cavity which is excavated by the flow. Such irregularities might originate from instabilities at the flow-cavity interface, as proposed by Liseau, Sandell & Knee (LSK88 (1988)), although there are no detailed numerical simulations of the process.
### 7.3. HH 11 and SVS 13
The interpretation for HH 11 and SVS 13 is more complicated because both sources are contained in the fields of view of the SWS and LWS instruments. Pre-shock densities $`n_010^5`$ cm<sup>-3</sup> are here found for the C component; this is a factor ten higher than in the other positions of the flow, which is not surprising given the close proximity of the origin of the flow. In these conditions the magnitude of the magnetic field (Eq. 1) could range from $`300`$$`\mu `$G to $``$1mG. Such high $`B_0`$ values have been claimed by Hartigan, Curiel & Raymond (HCR89 (1989)) to justify the relative faintness of HH 11 in 2.12$`\mu `$m H<sub>2</sub> images (see also Fig. 1). The possibility that FIR line emission toward SVS 13 arises in a collapsing envelope around is not relevant here because the predicted \[Oi\]63$`\mu `$m line flux (Ceccarelli, Hollenbach & Tielens CHT96 (1996)) is about 30 times lower than actually observed, while CO and H<sub>2</sub>O lines are below the detection limit for the present observations. Finally, there is the possibility that a fraction of the line fluxes measured with the LWS originates from the recently discovered embedded outflow source SVS 13B (Bachiller et al. Betal98 (1998)).
### 7.4. The Red-shifted Lobe
\[Oi\]63$`\mu `$m, \[Oi\]145$`\mu `$m and \[Siii\]34.8$`\mu `$m, together with a complement of H<sub>2</sub> and CO lines, have been detected toward the receding lobe, and define shock conditions which are not dramatically different from those present on the blue lobe. It is well known that no optical emission is detected toward the red lobe, and also published images in the H<sub>2</sub> 2.12$`\mu `$m line (Fig. 1, Garden et al. Getal90 (1990), Hodapp & Ladd HL95 (1995)) clearly show fainter emission there. Higher values of dust extinction with respect to the blue lobe have been invoked as an explanation for this asymmetry. Our observations, which trace similar shock conditions for the two lobes, tend to support this possibility.
## 8. The Shocks and the Molecular Outflow
If J-shocks between stellar winds and ambient material are responsible for the acceleration of the molecular outflow, a correlation (Hollenbach H85 (1985)) is expected between the outflow mass loss rate and the flux of the \[Oi\]63$`\mu `$m line, which is the dominant coolant in such shocks. The predicted mass loss rate, based on the observed \[Oi\]63$`\mu `$m cooling from HH 7, is $`4.8\times 10^6`$ $`M_{}`$ yr<sup>-1</sup> (see also Cohen et al. Cetal88 (1988), Ceccarelli et al. Cetal97 (1997)), in good agreement with the mass loss rate estimated by Lizano et al. (Letal88 (1988), see also Rodriguez et al. Retal90 (1990)) for the fast Hi wind believed to be responsible for the acceleration of the slow CO outflow (Snell & Edwards SE81 (1981)) . This fast neutral wind was also confirmed with CO observations by Bachiller & Cernicharo (BC90 (1990)).
Assuming momentum balance at the interface between the the wind and the ambient medium, Davis & Eislรถffel (DE95 (1995), DE96 (1996)) derived a simple relationship between the mechanical power of the wind $`L_w`$ and the power radiated by the shock $`L_{rad}`$:
$$\frac{L_{rad}}{L_w}=\frac{v_s}{v_w}\left[1\frac{v_s}{v_w}\right]^2$$
(2)
Since the working surfaces where the winds impact the medium are traced by the J-shocks, we make the assumption $`L_{rad}L_{O\mathrm{i}}+L_{Si\mathrm{ii}}+L_{H_2rovib}`$. The total cooling due to the near-IR H<sub>2</sub> vibrational lines measured by Gredel (G96 (1996)) along the flow is $`5\times 10^3`$ $`L_{}`$; we assume a slit width of few arcsecs, so we will conservatively multiply the observed value by 10 to allow for the extension of the HH objects. We estimate the mechanical power of the Hi wind according to:
$$L_w=\frac{1}{2}\frac{M_wv_w^2}{\tau _{dyn}}$$
(3)
Using the Hi parameters from Lizano et al. (Letal88 (1988)), we obtain $`L_w1.9`$ $`L_{}`$. Eq. (2) then provides $`v_s/v_w0.6`$, or $`v_{s_J}`$$`36`$ km s<sup>-1</sup> for an average Hi wind velocity of $`v_w`$60 km s<sup>-1</sup> (Lizano et al. Letal88 (1988)), in excellent agreement with our results.
## 9. Summary
The HH 7-11 flow, together with its red-shifted counterpart and SVS 13 (the candidate exciting source) have been studied via atomic, ionic and molecular spectroscopy. A complex scenario emerges, where:
1. we have detected atomic (\[Oi\]63$`\mu `$m, \[Oi\]145$`\mu `$m), ionic (\[Cii\]158$`\mu `$m, \[Siii\]34.8$`\mu `$m) and molecular (H<sub>2</sub>, CO and H<sub>2</sub>O) lines along the flow (both lobes) and toward SVS 13.
2. the low-excitation shock nature of the HH nebulosities along the flow is confirmed. Spectral signatures of C and J shocks are ubiquitously found along the HH 7-11 flow and its red-shifted counterpart. Our estimates for the shock velocities are $`v_{s_J}`$$`<\mathrm{\hspace{0.33em}40}50`$ km s<sup>-1</sup> and 15 $``$$`<`$$`v_{s_C}`$$``$$`<`$ 20 km s<sup>-1</sup>. The pre-shock density is $`10^4`$ cm<sup>-3</sup> toward the blue and the red lobe; for the C component only, we find $`n_010^5`$ cm<sup>-3</sup> at the location of SVS 13.
3. there is indirect evidence for an ordered B field oriented parallel to the direction of the flow. The magnitude of the magnetic field is $`B_0`$$`100`$$`\mu `$G on the lobes, increasing to $`300`$$`\mu `$G at the position of the flow origin; these figures, however, can vary of a factor 3 in each direction.
4. the gas-phase in the post C-shock region is deficient in H<sub>2</sub>O. We presented evidence that this may be due to freezing onto warm grains processed through the J-shock front and traveling downstream along the magnetic field lines.
5. the asymmetry in optical and NIR properties among the two lobes of the outflow is probably not caused by different pre-shock densities or shock velocities, supporting the hypothesis of higher extinction values toward the red lobe.
6. the total J-shock cooling is compatible with the molecular outflow being accelerated by the fast neutral wind detected in Hi and CO.
7. the whole flow area appears to be associated with a faint PDR illuminated by BD +30 549, the source responsible for the illumination of the whole NGC 1333 nebula.
We thank L. Testi and M. Cecere for their assistance with the observations and data reduction of the H<sub>2</sub> 2.12$`\mu `$m image presented in Fig. 1. The staff of the Mt. Palomar 60<sup>โฒโฒ</sup> telescope is also acknowledged. We also thank an anonymous referee whose comments improved the paper, and L.F. Rodriguez for his comments on an early version of this manuscript. The ISO Spectral Analysis Package (ISAP) is a joint development by the LWS and SWS Instrument Teams and Data Centers. Contributing institutes are CESR, IAS, IPAC, MPE, RAL and SRON. |
warning/0003/cond-mat0003218.html | ar5iv | text | # Turbo codes: the phase transition.
## 1 Introduction.
Communication through a noisy channel is a central problem in Information Theory . Error correcting codes are a widespread method for compensating the information corruption due to the noise, by cleverly increasing the redundancy of the message. Turbo codes are a recently invented class of error correcting codes with nearly optimal performances. They allows reliable communication (i.e. very low error per bit probability) with practical communication rates.
It is known, since the work of Sourlas , that there exists a close relationship between the statistical behavior of error correcting codes and the physics of some disordered spin models. Recently the tools developed in statistical physics have been employed in studying Gallager-type codes .
In Ref. the equivalence discovered by Sourlas is extended to turbo codes, and the basic features of the corresponding spin models are outlined. A remarkable property of a large family of turbo codes, presented in Ref. , is the existence of a no-error phase. In other words the error probability per bit vanishes beyond some critical (finite) signal to noise ratio. In Ref. some intuitive arguments supporting this thesis are given. Some analytical results concerning the critical value of the signal to noise ratio are announced without giving any derivation. These results are compared with numerical simulations.
In this paper we present the analytical results in their full generality, and explain their derivation. We prove the existence of the no-error phase and find the condition for its local stability. This condition is derived in two different approaches. In the first one we study the asymptotic dynamics of the decoding algorithm. In the second approach we use replicas and establish the condition for stability in the full replica space. Local stability is a necessary but not sufficient condition for the stability of the no-error phase. The critical signal to noise ratio obtained from local stability is the correct one only if the phase transition is a second order one: in the general case it is only a lower bound.
The spin models which are equivalent to turbo codes have the following statistical weight :
$`๐ซ(๐^{(1)},๐^{(2)}|๐ฑ,\beta )`$ $``$ $`{\displaystyle \frac{1}{Z(๐ฑ,\beta )}}{\displaystyle \underset{i=1}{\overset{N}{}}}\delta (ฯต_{\rho (i)}(๐^{(1)}),ฯต_i(๐^{(2)}))e^{\beta _{k=1}^2H^{(k)}(๐^{(k)})}`$ (1.1)
$`H^{(k)}(๐)`$ $``$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}J_i^{(k)}ฯต_i(๐){\displaystyle \underset{i=1}{\overset{N}{}}}h_i^{(k)}\eta _i(๐)`$ (1.2)
The dynamical variables of the model are the spins $`๐^{(k)}\{\sigma _1^{(k)},\mathrm{},\sigma _N^{(k)}\}`$ with $`k=1,2`$. We shall choose them to be Ising spins <sup>2</sup><sup>2</sup>2This corresponds to considering codes which works with a binary alphabet., that is $`\sigma _i^{(k)}=\pm 1`$. The spins enters in the hamiltonians $`H^{(k)}(๐)`$ through the local interaction terms $`ฯต_i(๐)`$ and $`\eta _i(๐)`$ which are products of $`\sigma `$โs. Their exact form can be encoded in two set of numbers $`\kappa (j;1)=0,1`$ and $`\kappa (j;2)=0,1`$ as follows: $`ฯต_i(๐)_{j=0}^r\sigma _{ij}^{\kappa (j;1)}`$ and $`\eta _i(๐)_{j=0}^r\sigma _{ij}^{\kappa (j;2)}`$. In order to fix completely our notation we set $`\kappa (0;1)=\kappa (0;2)=1`$. The quenched variables are:
* the couplings $`๐ฑ\{J_i^{(k)};h_i^{(k)}\}`$, whose distribution $`๐ซ(๐ฑ)_{i,k}P(J_i^{(k)})P(h_i^{(k)})`$ satisfies the conditions $`๐J_i^{(k)}P(J_i^{(k)})J_i^{(k)}>0`$ and $`๐h_i^{(k)}P(h_i^{(k)})h_i^{(k)}>0`$;
* the permutation $`\rho :\{1,\mathrm{},N\}\{1,\mathrm{},N\}`$, which has uniform distribution.
It is convenient to impose a fixed boundary condition at one end of the chain (i.e. $`\sigma _i=+1`$ for $`i0`$) and a free boundary condition at the other end. The model is composed by two one dimensional substructures (chains), which interact through the Kronecker delta functions in Eq. (1.1). When the average over permutations is taken into account this interaction turns into a mean field one. This interplay between the two subsystems, each one possessing a one dimensional structure, and the mean field interaction which couples them is clearly displayed by the analytical calculations. For further explanations on Eqs. (1.1-1.2) and their motivation we refer to .
The paper is organised as follows. In Sections 2 and 3 we present a first derivation of the stability condition. We write a โmean fieldโ equation which describes the dynamics of the decoding algorithm (Sec. 2), we show that it possesses a no-error fixed point and then study its behavior in a neighbourhood of this fixed point (Sec. 3). Thanks to this derivation we will understand how this fixed point is reached. In Section 4 replicas are introduced in order to compute the average over the permutations. We exhibit the no-error saddle point. In Section 5 the stability of the no-error saddle point is studied by diagonalizing the second derivative of the free energy. Finally in Section 6 the validity of our calculations is discussed. Appendix A collects some useful (although simple) facts of algebra. In Appendix B the type of integral equations which appear in Section 3 is studied in detail.
## 2 A โmean fieldโ equation for the decoding algorithm.
Some properties concerning the models defined by Eqs. (1.1-1.2) can be obtained by considering the โturbo decodingโ algorithm and making some factorization hypothesis. These hypothesis enable us to obtain a recursive integral equation for the probability distribution of a local field. They can be justified on heuristic grounds and arguments of this kind will be given later in this Section. Moreover the replica calculation presented in the Section 4 does support our arguments. In particular this approach allows us to derive the critical noise below which โperfectโ decoding is possible.
Turbo decoding is an iterative algorithm. The iteration variables are the fields $`๐ช^{(k)}\{\mathrm{\Gamma }_1^{(k)},\mathrm{},\mathrm{\Gamma }_N^{(k)}\}`$ with $`k=1,2`$. The step $`t`$ of the turbo decoding algorithm is defined as follows :
$`\mathrm{\Gamma }_i^{(1)}(t+1)`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}\text{arctanh}\left[ฯต_{\rho ^1(i)}(๐)_{\mathrm{\Gamma }^{(2)}(t)}^{(2)}\right]\mathrm{\Gamma }_{\rho ^1(i)}^{(2)}(t)`$ (2.1)
$`\mathrm{\Gamma }_i^{(2)}(t+1)`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}\text{arctanh}\left[ฯต_{\rho (i)}(๐)_{\mathrm{\Gamma }^{(1)}(t)}^{(1)}\right]\mathrm{\Gamma }_{\rho (i)}^{(1)}(t)`$ (2.2)
The expectation value $`_{\mathrm{\Gamma }^{(k)}}^{(k)}`$ is intended to be taken with respect to the Boltzmann weight with the modified hamiltonian $`H^{(k)}(๐)_{i=1}^N\mathrm{\Gamma }_i^{(k)}ฯต_i(๐)`$. The iteration variables $`\mathrm{\Gamma }_i^{(k)}`$ should be interpreted as external fields conjugate to the operators $`ฯต_i(๐^{(k)})`$. They describe, in an approximate way, the action of each of the two chains on the other one.
In order to lighten the notation, let us write Eqs. (2.1-2.2) in the form:
$`๐ช^{(k)}(t+1)=F_\rho ^{(k)}(๐ช^{(k^{})}(t),๐ฑ^{(k^{})})`$ (2.3)
with $`k^{}=2`$ if $`k=1`$, and $`k^{}=1`$ if $`k=2`$. Due to the randomness in the couplings $`๐ฑ`$, the fields $`๐ช`$ are random variables. Equation (2.3) implies an integral equation for the probability distribution of $`๐ช`$:
$`๐ซ_{t+1}(๐ช^{(k)})={\displaystyle ๐๐ช^{(k^{})}๐๐ฑ^{(k^{})}๐ซ_t(๐ช^{(k^{})},๐ฑ^{(k^{})})\delta \left[๐ช^{(k)}F_\rho ^{(k)}(๐ช^{(k^{})},๐ฑ^{(k^{})})\right]}`$ (2.4)
Let us state a few approximations which allow us to reduce Eq. (2.4) to a much simpler one.
* We make the substitution $`๐ซ_t(๐ช^{(k^{})},๐ฑ^{(k^{})})๐ซ_t(๐ช^{(k^{})})๐ซ(๐ฑ^{(k^{})})`$ in Eq. (2.4). This yields a closed integral equation describing the evolution of the distribution $`๐ซ_t(๐ช^{(k)})`$.
* We neglect correlations between the fields at different sites:
$`๐ซ_t(๐ช^{(k)}){\displaystyle \underset{i=1}{\overset{N}{}}}\pi _{i,t}^{(k)}(\mathrm{\Gamma }_i^{(k)})`$ (2.5)
These two hypothesis imply that Eq. (2.4) is equivalent to:
$`\pi _{i,t+1}^{(k)}(y)`$ $`=`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}๐\pi _t^{(k^{})}[๐]{\displaystyle ๐๐ซ[๐ฑ]\delta \left(y\frac{1}{\beta }\text{arctanh}\left(ฯต_{\widehat{\rho }(i)}(๐)_{๐ฑ,๐}\right)+x_{\widehat{\rho }(i)}\right)}`$ (2.6)
$`d\pi _t^{(k^{})}[๐]`$ $``$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}dx_i\pi _t^{(k^{})}(x_i)`$ (2.7)
where $`\widehat{\rho }`$ is the appropriate permutation of $`\{1,\mathrm{},N\}`$, i.e. $`\widehat{\rho }=\rho ^1`$ if $`k=1`$ and $`\widehat{\rho }=\rho `$ if $`k=2`$. The expectation value $`_{๐ฑ,๐}`$ on the right hand side of Eq. (2.6) has to be taken with respect to the hamiltonian $`H(๐)_{i=1}^N(J_i+x_i)ฯต_i(๐)_{i=1}^Nh_i\eta _i(๐)`$.
Let us now define a field distribution averaged over the permutations and the sites:
$`\pi _t^{(k)}(x){\displaystyle \frac{1}{N!}}{\displaystyle \underset{\rho }{}}{\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}\pi _{i,t}^{(k)}(x|\rho )`$ (2.8)
where we made explicit the dependence of $`\pi _{i,t}^{(k)}`$ upon the specific permutation $`\rho `$ which defines the code. We can now state our last approximation.
* We make the substitution $`\pi _{i,t}^{(k)}(x|\rho )\pi _t^{(k)}(x)`$ on the right hand side of Eq. (2.6).
This yields a recursive equation for $`\pi _t^{(k)}`$:
$`\pi _{t+1}(y)={\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐\pi _t[๐]{\displaystyle ๐๐ซ[๐ฑ]\delta \left(y\frac{1}{\beta }\text{arctanh}\left(ฯต_i(๐)_{๐ฑ,๐}\right)+x_i\right)}`$ (2.9)
The indices $`(k)`$ and $`(k^{})`$ have been dropped since we can define $`\pi _t=\pi _t^{(1)}`$ for $`t`$ odd, and $`\pi _t=\pi _t^{(2)}`$ for $`t`$ even, or vice-versa. A byproduct of this heuristic derivation is the expression for the probability distribution of the expectation values $`ฯต_i(๐)`$ after $`t`$ iterations of the turbo decoding algorithm: $`๐ซ_t(ฯต)=\frac{1}{N}_{i=1}^N_{\mathrm{}}^+\mathrm{}๐\pi _t[๐]๐๐ซ[๐ฑ]\delta \left(ฯตฯต_i(๐)_{๐ฑ,๐}\right)`$.
Let us discuss the validity of the approximations made in deriving Eq. (2.9).
* These approximations should be accurate in the thermodynamic limit for a generic random permutation $`\rho `$. The reason is that the correlations produced by Eqs. (2.1-2.2) have short range: $`ฯต_i(๐)`$ and $`ฯต_j(๐)`$ have a significant correlation only if $`|ij|`$ is less than some characteristic length. The random permutation $`\rho `$ reshuffles the sites so that the correlation between two fields $`\mathrm{\Gamma }_i^{(k)}`$ and $`\mathrm{\Gamma }_j^{(k)}`$ is vanishing with high probability if $`|ij|`$ is required to be โsmallโ. The correlations which โsurviveโ (non vanishing only between โdistantโ sites) are irrelevant when computing the expectation values of local operators. In order to make this last assertion plausible, let us suppose that, for each site $`i`$, we can find a โlargeโ <sup>3</sup><sup>3</sup>3Here โlargeโ means that $`lim_N\mathrm{}L(N)=\mathrm{}`$. interval $`[iL(N),i+L(N)]`$ of the chain, such that the correlations between the couplings inside the interval are negligible. The expectation value $`ฯต_i(๐)_{๐ฑ,๐}`$ will not depend (as $`N\mathrm{}`$) upon the couplings outside $`[iL(N),i+L(N)]`$ (this is always true in one dimension at non zero temperature) and can be then safely computed without taking into account the correlations. It is easy to find a similar argument concerning the correlations between $`๐ช^{(k^{})}`$ and $`๐ฑ^{(k^{})}`$ in Eq. (2.4).
* This is the probabilistic analogue of the replica symmetric approximation. Let us consider the fixed point equation $`\pi _{t+1}=\pi _t`$ corresponding to the dynamics defined by Eq. (2.9). It is remarkable that this fixed point equation coincides with the saddle point equation obtained by the standard replica method in the replica symmetric approximation (see Section 4). This fact confirms our conclusions about the relevance of the various approximations.
## 3 The behavior of the decoding algorithm.
Equation (2.9) is the final outcome of our heuristic derivation. We want to study its behavior when the distribution $`\pi (x)`$ is concentrated on large values of the field $`x`$, that is when the error probability is very small. In this regime the most relevant spin configuration satisfies $`ฯต_i(๐)=+1`$ for each $`i=1,\mathrm{},N`$. The lowest excitations are such that $`ฯต_i(๐)=1`$ only on a few sites. The first crucial point will be to understand that, for a class of hamiltonians of the type (1.2) (which will be defined as โrecursive โ), the energy to be paid for flipping a single $`ฯต`$ variable diverges in the thermodynamic limit. The second point will be to evaluate the energy to be paid for flipping two $`ฯต`$ variables. In order to treat both these passages in full generality it is convenient to use an algebraic bookkeeping technique which we shall soon explain. The results concerning these two points will be useful again in Section 5.
A preliminary step consists in making the change of variables $`X_ie^{2\beta x_i}`$ and introducing the corresponding distribution function $`Q_t(X)dX=\pi _t(x)dx`$. Low $`X`$โs correspond then to large local fields, i.e. to low error probability. The result is
$`Q_{t+1}(Y)={\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^{\mathrm{}}}๐Q_t[๐ฟ]{\displaystyle ๐๐ซ[๐ฑ]\delta \left(Y\frac{1}{X_i}\frac{Z(ฯต_i(๐)=1;๐ฑ,๐ฟ)}{Z(ฯต_i(๐)=+1;๐ฑ,๐ฟ)}\right)}`$ (3.1)
where
$`Z(ฯต_i(๐)=ฯต;๐ฑ,๐ฟ)Z_i(ฯต)={\displaystyle \underset{๐:ฯต_i(๐)=ฯต}{}}e^{\beta H(๐)}{\displaystyle \underset{k=1}{\overset{N}{}}}X_k^{\frac{1}{2}(1ฯต_k(๐))}`$ (3.2)
with $`H(๐)=_iJ_iฯต_i(๐)_ih_i\eta _i(๐)`$. Let us introduce some notations in order to write down the small $`X`$ expansion of $`Z_i(ฯต)`$: $`(k_1,\mathrm{},k_l)`$ is an $`l`$-uple (not ordered) of integers in $`\{1,\mathrm{},i1,i+1,\mathrm{},N\}`$; $`๐_0`$ is the configuration such that $`ฯต_i(๐)=+1`$ for all the sites $`i`$; $`๐(k,l,m,\mathrm{})`$ is the configuration such that $`ฯต_j(๐)=1`$ if $`j=k,l,m,\mathrm{}`$ and $`ฯต_j(๐)=1`$ otherwise (there is at most one such configuration once the boundary conditions have been specified); $`E_0H(๐_0)`$ is the energy of the ordered configuration; finally $`\mathrm{\Delta }(k,l,m,\mathrm{})H(๐(k,l,m,\mathrm{}))H(๐_0)`$. The following expressions are straightforward:
$`Z_i(+1)=e^{\beta E_0}{\displaystyle \underset{l=0}{\overset{N1}{}}}{\displaystyle \underset{(k_1,\mathrm{},k_l)}{}}X_{k_1}\mathrm{}X_{k_l}e^{\beta \mathrm{\Delta }(k_1,\mathrm{},k_l)}`$ (3.3)
$`Z_i(1)=X_ie^{\beta E_0}{\displaystyle \underset{l=0}{\overset{N1}{}}}{\displaystyle \underset{(k_1,\mathrm{},k_l)}{}}X_{k_1}\mathrm{}X_{k_l}e^{\beta \mathrm{\Delta }(i,k_1,\mathrm{},k_l)}`$ (3.4)
The โbookkeeping technique โ which we shall adopt in treating the above expansions consists in using the algebra of โgenerating polynomialsโ . This approach allows us to consider a general hamiltonian of the type (1.2). Let us define the following polynomials on $`_2`$: $`G(x)_{j=1}^{\mathrm{}}G_jx^j`$, with $`\sigma _j=(1)^{G_j}`$; $`g_n(x)=_{j=0}^r\kappa (j;n)x^j`$; $`๐ข^{(n)}(x)g_n(x)G(x)_{j=1}^{\mathrm{}}๐ข_j^{(n)}x^j`$. Notice that the boundary condition on $`๐`$ can be translated as follows: $`G(x)`$ is a series of strictly positive powers of $`x`$.
It is necessary to distinguish two types of models: in the first case $`g_1(x)`$ divides $`g_2(x)`$, i.e. $`g_2(x)/g_1(x)`$ is a polynomial (these are the โnon recursiveโ models, a particular case being $`ฯต_i(๐)=\sigma _i`$); in the second one $`g_1(x)`$ does not divide $`g_2(x)`$, i.e. $`g_2(x)/g_1(x)`$ is a series (โrecursiveโ models).
We shall treat the โrecursiveโ models first. In this case the first order terms in the expansions (3.3) and (3.4) are exponentially small in the size. In order to prove this assertion, let us consider the configuration $`๐(l)`$. The relevant generating polynomials are $`๐ข^{(1)}(x)=x^l`$ and $`๐ข^{(2)}(x)=x^lg_2(x)/g_1(x)`$. The form of $`๐ข^{(2)}(x)`$ is given by the following result of algebra
###### Lemma 3.1
Let $`g(x)`$ and $`f(x)`$ be two polynomials on $`_2`$ such that $`g(0)=f(0)=1`$, $`f(x)1`$, and their greatest common divisor $`\mathrm{gcd}(f(x),g(x))`$ is equal to 1. Then there exists an integer $`\omega `$ such that $`g(x)/f(x)=_{n=0}^{\mathrm{}}x^{n\omega }p_n(x)`$ with $`\mathrm{deg}[p_n(x)]<\omega `$ and $`p_n(x)=p_{\mathrm{}}(x)0`$ if $`n`$ is large enough. Hereafter we shall call $`\omega (f)`$ the smallest of such integers.
An explicit expression for $`\omega (f)`$ is given in the Appendix A. The Lemma 3.1 applies to our case if we divide both $`g_1(x)`$ and $`g_2(x)`$ by their greater common divisor: $`f_k(x)g_k(x)/\mathrm{gcd}(g_1(x),g_2(x))`$, so that $`\mathrm{gcd}(f_1(x),f_2(x))=1`$. It implies that if we write down the numbers $`\eta _j(๐(i))=\pm 1`$ we get an antiperiod followed by a non trivial periodic sequence with period $`\omega (f_1)`$. Let us consider a site โin the bulkโ: $`N\delta <i<N(1\delta )`$ with $`\delta `$ a (small) positive number. Then, using the convention $`h_j=0`$ for $`j>N`$, we get:
$`\mathrm{\Delta }(i)`$ $`=`$ $`2J_i+2{\displaystyle \underset{j=1}{\overset{N}{}}}๐ข_j^{(2)}h_j=2J_i+2{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{k=0}{\overset{\omega (f_1)1}{}}}p_{n,k}h_{i+n\omega (f_1)+k}`$ (3.5)
which diverges almost surely in the thermodynamic limit if $`h>0`$ (see the Introduction on this point). In Eq. (3.1) we must sum also terms which are โnearโ the boundaries, i.e. $`iN\delta `$ or $`iN(1\delta )`$. These give however a negligible contribution.
Let us now consider the second order terms of the expansions (3.3) and (3.4). They involve configurations $`๐(k,l)`$ with two flipped $`ฯต(๐)`$โs. The only configurations which give a non negligible contribution are the ones which involve a finite (in the $`N\mathrm{}`$ limit) number of flipped $`\eta (๐)`$โs. This corresponds to choosing $`k`$ and $`l`$ such that $`(x^k+x^l)g_2(x)/g_1(x)`$ is a polynomial (and not an infinite series). The following useful result is proved in the Appendix A.
###### Lemma 3.2
Let $`f(x)`$ be a polynomial on $`_2`$ such that $`f(0)=1`$ and $`k`$ an integer. Then there exists an integer $`\omega (f)`$ such that $`f(x)`$ divides $`1+x^k`$ if and only if $`k`$ is a strictly positive multiple of $`\omega (f)`$.
As suggested by the notation the $`\omega (f)`$โs cited in Lemmas 3.1 and 3.2 are indeed equal. The terms which give a non vanishing contribution at order $`X^2`$ in the expansions (3.3-3.4) are the ones corresponding to configurations $`๐(k,l)`$ such that $`|kl|`$ is a multiple of $`\omega (f_1)`$. In order to evaluate these terms we must count the number of flipped $`\eta (๐)`$โs. This number is nothing but the number of non zero coefficients in the polynomial $`(x^k+x^l)g_2(x)/g_1(x)`$. Let us define the weight of a polynomial $`p(x)=_kp_kx^k`$ over $`_2`$ as the number of its non zero coefficients: $`\text{weight}(p)\mathrm{\#}\{p_k|p_k0\}`$. The weight of $`(x^k+x^l)g_2(x)/g_1(x)`$ is given, for a large class of hamiltonians, by the following lemma.
###### Lemma 3.3
Let $`f(x)`$ and $`g(x)`$ be two polynomial on $`_2`$ such that $`f(0)=g(0)=1`$, $`f(x)1`$, and $`\mathrm{gcd}(f(x),g(x))=1`$. If $`\mathrm{deg}(g)\omega (f)`$ then the weight of $`s_m(x)(1+x^{m\omega (f)})g(x)/f(x)`$ is given by $`\mathrm{weight}(s_m)=w_0(f,g)+w_1(f,g)m`$ for each $`m1`$. The coefficients $`w_0(f,g)`$ and $`w_1(f,g)`$ are positive integers whose explicit expressions are given by Eqs. (A.8-A.9).
Appendix A contains also an illustration of what could happen in the more general case.
By using Lemmas 3.2 and 3.3 we can linearize with respect to $`X`$ the expression on the r.h.s. of Eq. (3.1):
$`{\displaystyle \frac{1}{X_i}}{\displaystyle \frac{Z_i(1)}{Z_i(+1)}}={\displaystyle \underset{m0}{}}X_{i+m\omega (f_1)}e^{\beta \mathrm{\Delta }(i,i+m\omega (f_1))}+O(X^2)`$ (3.6)
and defining $`s_m(x)(1+x^{m\omega (f_1)})g_2(x)/g_1(x)=_js_{m,j}x^j`$ we get
$`\mathrm{\Delta }(k,l)=2J_k+2J_l+2{\displaystyle \underset{j}{}}s_{m,j}h_{\mathrm{min}(k,l)+j}`$ (3.7)
if $`|kl|=m\omega (f_1)`$. Clearly Eq. (3.6) holds only for $`i`$ in the โbulkโ (i.e. $`N\delta <i<(1\delta )N`$) up to terms which are exponentially small in the size $`N`$.
Our first important observation is that the right hand side of Eq. (3.6) vanishes if $`X_k=0`$ for $`k=1,\mathrm{},N`$. This means that $`Q_{}(X)=\delta (X)`$ is a fixed point of Eq. (3.1) for โrecursiveโ models. Recall that the change of variables which yields Eq. (3.1) is $`X=e^{2\beta x}`$ and that $`x`$ has the meaning of an effective field acting on $`ฯต_i(๐)`$. The solution $`Q_{}(X)`$ corresponds then to a phase with completely frozen spins: $`ฯต_i(๐)=+1`$.
We would like to understand if this phase is stable for some temperature $`\beta `$ and some distribution of the couplings. A possible approach is to study the turbo decoding dynamics (as described by Eq. (3.1)) when starting from a distribution โnearโ $`Q_{}(X)`$. Let us suppose that, for $`Q_t(X)`$ near enough to $`Q_{}(X)`$, we can safely neglect $`O(X^2)`$ terms on the r.h.s. of Eq. (3.6):
$`Q_{t+1}(Y)={\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^{\mathrm{}}}๐Q_t[๐ฟ]{\displaystyle ๐๐ซ[๐ฑ]\delta \left(Y\underset{m0}{}X_{i+m\omega (f_1)}e^{\beta \mathrm{\Delta }(i,i+m\omega (f_1))}\right)}`$ (3.8)
This equation is very similar to a class of recursive equations which appear in a completely different context: polymers on disordered trees . These are of the type
$`P_{t+1}(Z)={\displaystyle _0^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{K}{}}}dZ_iP_t(Z_i){\displaystyle \rho (V_1,\mathrm{},V_K)๐V_1\mathrm{}๐V_K\delta \left(Z\underset{i=1}{\overset{K}{}}e^{\beta V_i}Z_i\right)}`$ (3.9)
The only non trivial difference is that the linear function of $`๐ฟ`$ appearing inside the delta function on the r.h.s. of Eq. (3.8) depends upon a macroscopic (indeed linear in $`N`$) number of $`X`$โs. In Eq. (3.9), instead, only a finite number of variables appears: $`K`$ is the coordination number of the tree minus one. Notice however that, for $`m`$ large, $`\mathrm{\Delta }(i,i+m\omega (f_1))2\text{weight}(s_m)h2w_1(f_1,f_2)mh`$. We can thus truncate the sum in Eq. (3.8) to $`mM`$ by making an error of order $`O(e^{cM})`$ and we guess that the limit $`M\mathrm{}`$ can be taken at the end without problems <sup>4</sup><sup>4</sup>4This argument is not mathematically rigorous since it is not honest to use the central limit theorem in this case: we refer to Appendix B for more convincing arguments..
Let us summarize some results of which are useful in our discussion. It turns out that Eq. (3.9) is equivalent to a discretization of the Kolmogorov-Petrovsky-Piscounov (KPP) equation (a well studied partial differential equation). Using this equivalence the large time limit of Eq. (3.9) is obtained:
$`P_t(X)e^{\beta c(\beta )t}\overline{P}(Xe^{\beta c(\beta )t})`$ (3.10)
corresponding to a front wave solution of the KPP equation with front velocity $`c(\beta )`$. If we define the function
$`v(\beta ){\displaystyle \frac{1}{\beta }}\mathrm{log}\left({\displaystyle \underset{i=1}{\overset{K}{}}}{\displaystyle ๐V_1\mathrm{}๐V_K\rho (V_1,\mathrm{},V_K)e^{\beta V_i}}\right)`$ (3.11)
then the front velocity is given by the following construction:
$`c(\beta )=\{\begin{array}{ccc}v(\beta )& \text{if}& \beta \beta _c\\ v(\beta _c)& \text{if}& \beta >\beta _c\end{array}`$ (3.14)
with $`\beta _c`$ given by
$`{\displaystyle \frac{d}{d\beta }}|_{\beta _c}v(\beta )=0`$ (3.15)
At the critical temperature $`\beta _c`$ a freezing phenomenon takes place with the front velocity sticking to its minimal value.
Let us apply these results to our case, i.e. to Eq. (3.8). The large time solution $`Q_t(X)e^{\beta c(\beta )t}\overline{Q}(Xe^{\beta c(\beta )t})`$ gives the correct behavior for $`t\mathrm{}`$ only if $`c(\beta )<0`$. In this case $`lim_t\mathrm{}Q_t(X)=Q_{}(X)`$ and it is then correct to linearize Eq. (3.1): the frozen phase is stable. If, on the other hand, $`c(\beta )0`$ then we must take into account higher order terms in the low $`X`$ expansion and the asymptotic form is no longer of the type defined by Eq. (3.10): the frozen phase is unstable.
In the thermodynamic limit we get
$`e^{\beta v(\beta )}`$ $`=`$ $`{\displaystyle \underset{m0}{}}{\displaystyle ๐๐ซ[๐ฑ]e^{\beta \mathrm{\Delta }(i,i+m\omega (f_1))}}=`$
$`=`$ $`2\left({\displaystyle ๐JP(J)e^{2\beta J}}\right)^2{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left({\displaystyle ๐hP(h)e^{2\beta h}}\right)^{\mathrm{weight}(s_m)}`$
The front velocity $`c(\beta )`$ is obtained by applying the construction given in Eqs. (3.14-3.15) to Eq. (3). If the hypothesis of Lemma 3.3 are satisfied we can easily sum the series:
$`e^{\beta v(\beta )}`$ $`={\displaystyle \frac{2\left({\displaystyle ๐JP(J)e^{2\beta J}}\right)^2\left({\displaystyle ๐hP(h)e^{2\beta h}}\right)^{w_0(f_1,f_2)+w_1(f_1,f_2)}}{1\left({\displaystyle ๐hP(h)e^{2\beta h}}\right)^{w_1(f_1,f_2)}}}`$ (3.17)
We discuss now Eq. (3.17), the more general case being completely analogous. The series converges only if $`๐hP(h)e^{2\beta h}<1`$. If $`๐hP(h)h>0`$, as we supposed sinc the beginning, then convergence is assured for $`0<\beta <\beta _1`$ with $`๐hP(h)e^{2\beta _1h}=1`$. It is easy to see that $`\beta v(\beta )`$ is strictly convex for $`0<\beta <\beta _1`$ and thus $`v(\beta )`$ has either one global minimum or is strictly monotonic for $`0<\beta <\beta _1`$. Since $`lim_{\beta 0^+}v(\beta )=lim_{\beta \beta _1^{}}v(\beta )=+\mathrm{}`$ the first possibility is excluded and we conclude that $`0<\beta _c<\beta _1`$. The important point is that the right hand side of Eq. (3) is well defined every time we need of it, i.e. for $`0<\beta <\beta _c`$.
In applications to turbo codes a simplification occurs: we are interested in a particular temperature, $`\beta =1`$, and we are left with a unique parameter: the signal to noise ratio $`1/w^2`$. Moreover the probability distributions of the couplings are fixed by the characteristics of the communication channel . If we introduce the auxiliary variables $`\widehat{J}`$ and $`\widehat{h}`$, which correspond to the output of the channel, the probability distributions are obtained as follows
$`P(J)dJ=P(\widehat{J}|+1)d\widehat{J}\text{with}J={\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{P(\widehat{J}|+1)}{P(\widehat{J}|1)}}`$ (3.18)
where $`P(\widehat{J}|\tau )`$ is the probability distribution of the output of the channel conditional to the input $`\tau `$. A similar expression holds for $`h`$. If the channel is symmetric (i.e. if $`P(\widehat{J}|1)=P(\widehat{J}|+1)`$) one easily obtains $`\beta _1=1`$ and then $`c(\beta =1,w^2)=v(\beta _c,w^2)`$. We can distinguish the two cases defined below.
* If $`v(\beta ,w^2)<0`$ for some $`0<\beta <1`$ then we are in the no-error phase and the turbo decoding algorithm converges to the message with โvelocityโ $`c(\beta =1,w^2)=\mathrm{min}_{0<\beta <1}v(\beta ,w^2)`$. We expect the condition $`v(\beta _c,w^2)<0`$ to be verified in the โlow noiseโ region $`w^2<w_{loc}^2`$.
* If $`v(\beta ,w^2)0`$ in the interval $`0<\beta <1`$ then $`c(\beta =1,w^2)0`$ and the linearization in Eq. (3.8) is no longer reliable. In this case $`\pi _t(x)`$ is expected to converge for $`t\mathrm{}`$ to some distribution supported on finite fields $`x`$. The decoded message will be plagued by a finite error probability per bit, no matter how many times do we iterate the turbo decoding algorithm.
Let us now study some examples. We consider a gaussian channel with:
$`P(\widehat{J}|\tau )={\displaystyle \frac{1}{(4\pi w^2)^{1/2}}}\mathrm{exp}\left\{{\displaystyle \frac{(\widehat{J}\tau )^2}{4w^2}}\right\}`$ (3.19)
$`P(\widehat{h}|\tau )={\displaystyle \frac{1}{(2\pi w^2)^{1/2}}}\mathrm{exp}\left\{{\displaystyle \frac{(\widehat{h}\tau )^2}{2w^2}}\right\}`$ (3.20)
This choice of the variances is justified since it corresponds to a code with rate $`1/3`$ (see Ref. ). It is useful to define the function
$`z(\beta ,w^2)={\displaystyle ๐hP(h)e^{2\beta h}}=\left({\displaystyle ๐JP(J)e^{2\beta J}}\right)^2=\mathrm{exp}\left[{\displaystyle \frac{2\beta (\beta 1)}{w^2}}\right]`$ (3.21)
The three cases below have been already considered in Ref. . We refer to the Appendix A for the calculation of the constants $`w_0`$ and $`w_1`$ to be used in Eq. (3.17).
1. A model with nearest neighbours interaction is: $`ฯต_i(๐)\sigma _i\sigma _{i1}`$ and $`\eta _i(๐)=\sigma _i`$ (which corresponds to the generating polynomials $`g_1(x)=1+x`$ and $`g_2(x)=1`$). Using Eq. (3.17) and the fact that $`w_0(f_1,f_2)=0`$ and $`w_1(f_1,f_2)=1`$ we get
$`v(\beta ,w^2)={\displaystyle \frac{1}{\beta }}\mathrm{log}{\displaystyle \frac{2z^2(\beta ,w^2)}{1z(\beta ,w^2)}}`$ (3.22)
It is easy to see that $`v(\beta ,w^2)0`$ for each $`0<\beta <1`$ if $`w^2w_{loc}^2=1/\mathrm{log}4`$.
2. For $`ฯต_i(๐)\sigma _i\sigma _{i1}\sigma _{i2}`$ and $`\eta _i(๐)=\sigma _i\sigma _{i2}`$ (generating polynomials: $`g_1(x)=1+x+x^2`$ and $`g_2(x)=1+x^2`$) we obtain $`w_0(f_1,f_2)=2`$ and $`w_1(f_1,f_2)=2`$ and then
$`v(\beta ,w^2)={\displaystyle \frac{1}{\beta }}\mathrm{log}{\displaystyle \frac{2z^5(\beta ,w^2)}{1z^2(\beta ,w^2)}}`$ (3.23)
Finally $`w_{loc}^2=1/(2\mathrm{log}z_c)`$ where $`z_c`$ is the only real solution of the equation $`2z^5+z^2=1`$.
3. If we consider the model given by $`ฯต_i(๐)\sigma _i\sigma _{i1}\sigma _{i2}\sigma _{i3}\sigma _{i4}`$ and $`\eta _i(๐)=\sigma _i\sigma _{i4}`$ (generating polynomials: $`g_1(x)=1+x+x^2+x^3+x^4`$ and $`g_2(x)=1+x^4`$) we obtain $`w_0(f_1,f_2)=2`$ and $`w_1(f_1,f_2)=2`$ as in the previous example. Both $`v(\beta ,w^2)`$ and $`w_{loc}^2`$ coincide with the ones obtained above.
Let us make a few observations about the validity of our calculation. The threshold $`w_{loc}^2`$ has been obtained by starting from a distribution $`Q(X)`$ very near to the โfrozenโ one $`Q_{}(X)`$ and linearizing Eq. (3.1) in $`X`$. It must then be interpreted as a threshold for local stability of the โfrozenโ solution. Moreover, if we take seriously the heuristic derivation of Eq. (2.9), we can deduce something about the dynamics of the turbo decoding algorithm in the error-free phase: the probability distribution of the auxiliary fields $`\mathrm{\Gamma }_i^{(k)}(t)`$ moves towards infinitely large fields with an average velocity $`c(\beta ,w^2)`$. This conclusion is compared with numerical data in Fig. (1): the agreement seems to be quite good. An interesting outcome of the previous calculation is that the approach to the perfect decoding becomes slower near to the critical signal to noise ratio.
Let us now discuss the โnon recursiveโ models, that is models such that $`g_1(x)`$ divides $`g_2(x)`$. In this case the energy $`\mathrm{\Delta }(i)`$ to be paid for flipping $`\eta _i(๐)`$ remains finite in the thermodynamic limit. The low $`X`$ expansions in Eqs. (3.3-3.4) have a non vanishing term of order $`O(X)`$. This implies that $`Q_{}(X)=\delta (X)`$ is no longer a fixed point of Eq. (3.1). Let us compute $`\mathrm{\Delta }(i)`$. For โnon recursiveโ models we can define the polynomial $`s(x)_ks_kx^kg_2(x)/g_1(x)`$. It is easy to show that $`\mathrm{\Delta }(i)=2J_i+2_ks_kh_{i+k}`$. A simple approximation of the fixed point of Eq. (3.1) is:
$`Q_{\mathrm{}}(X){\displaystyle ๐JP(J)\underset{i=1}{\overset{w}{}}dh_iP(h_i)\delta \left(Xe^{2\beta J2\beta _{i=1}^wh_i}\right)}`$ (3.24)
with $`w\mathrm{weight}(s)`$. This approximation is supposed to be good in the low noise region where we expect the distributions $`Q_t(X)`$ to be concentrated on small $`X`$โs.
## 4 The replica calculation.
The replica method starts with the computation of the (integer) moments of the partition function. This can be done by introducing an appropriate order parameter (the choice is a matter of convenience) and by recurring to standard tricks. Here we choose to use the (multi)-overlaps $`q_{a_1\mathrm{}a_l}`$ and their complex conjugates $`\widehat{q}_{a_1\mathrm{}a_l}`$:
$`\overline{Z^n}`$ $`=`$ $`{\displaystyle \frac{N}{\pi }๐q_0๐\widehat{q}_0\underset{a}{}\frac{N}{\pi }dq_ad\widehat{q}_a\underset{(a,b)}{}\frac{N}{\pi }dq_{ab}d\widehat{q}_{ab}\mathrm{}e^{NS[q,\widehat{q}]}}`$ (4.1)
$`S[q,\widehat{q}]`$ $`=`$ $`1+q_0\widehat{q}_0+{\displaystyle \underset{a}{}}q_a\widehat{q}_a+{\displaystyle \underset{(a,b)}{}}q_{ab}\widehat{q}_{ab}+\mathrm{}+n\mathrm{log}2+`$
$`+\beta _{1d,n}[q]+\beta _{1d,n}[\widehat{q}]`$
$`_{1d,n}[q]`$ $``$ $`\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{N\beta }}\mathrm{log}Z_{1d,n}[q]`$ (4.3)
$`Z_{1d,n}[q]`$ $``$ $`{\displaystyle \underset{\{\sigma _i^a\}}{}}{\displaystyle \underset{i=1}{\overset{N}{}}}[q_0+{\displaystyle \underset{a}{}}q_aฯต_i(๐^a)+{\displaystyle \underset{(a,b)}{}}q_{ab}ฯต_i(๐^a)ฯต_i(๐^b)+\mathrm{}]`$
$`{\displaystyle }d๐ซ[๐ฑ]\mathrm{exp}\{\beta {\displaystyle \underset{a}{}}H(๐^a;๐ฑ)\}`$
where $`H(๐;๐ฑ)=_iJ_iฯต_i(๐)_ih_i\eta _i(๐)`$. and the replica indices $`a,b,\mathrm{}`$ run from $`1`$ to $`n`$. The usual mean field models have no geometrical structure at all. In those cases the introduction of the order parameters leads to a (replicated) partition function which factorizes over the sites. In our case we are left with the problem of computing the one dimensional partition functions $`Z_{1d,n}[q]`$. These correspond to the one dimensional sub-structures which are not destroyed by the randomness of the model. The saddle point equations are easily written
$`\widehat{q}_{a_1\mathrm{}a_l}=\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{ฯต_i(๐^{a_1})\mathrm{}ฯต_i(๐^{a_l})}{[q_0+_aq_aฯต_i(๐^a)+\mathrm{}]}}_q`$ (4.5)
where the expectation values $`()_q`$, $`()_{\widehat{q}}`$ are defined as follows
$`()_q`$ $``$ $`{\displaystyle \frac{1}{Z_{1d,n}[q]}}{\displaystyle ๐๐ซ[๐ฑ]\underset{\{\sigma _i^a\}}{}()\underset{i=1}{\overset{N}{}}[q_0+\underset{a}{}q_aฯต_i(๐^a)+\mathrm{}]e^{\beta _aH[๐{}_{}{}^{a};๐ฑ]}}`$ (4.6)
In the recursive case Eq. (4.5) admits the following solution <sup>5</sup><sup>5</sup>5In fact there is a one parameter family of solutions which are degenerate. This fact is due to a (not very interesting) symmetry of the action (4): $`S[q,\widehat{q}]=S[e^{i\theta }q,e^{i\theta }\widehat{q}]`$. However the integration over the parameter $`\theta `$ does not pose any problem. We shall fix this freedom by imposing $`q_0`$ to be real. corresponding to a no-error phase: $`q_{a_1\mathrm{}a_l}^{}=\widehat{q}_{b_1\mathrm{}b_l}^{}=2^{n/2}`$. The free energy of this phase is $`f_0(\beta )=2๐JP(J)J2๐hP(h)h`$. If we parametrize the replica symmetric ansatz as in Ref.
$`q_{a_1\mathrm{}a_l}={\displaystyle _{\mathrm{}}^+\mathrm{}}๐x\pi (x)\mathrm{cosh}^n(\beta x)\mathrm{tanh}^l(\beta x)`$ (4.7)
and analogously for $`\widehat{q}_{b_1\mathrm{}b_m}`$ (with a different distribution $`\widehat{\pi }(x)`$), the following free energy functional can is obtained in the limit $`n0`$:
$`f[\pi ,\widehat{\pi }]`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle ๐x๐y\pi (x)\widehat{\pi }(y)\mathrm{log}\left[2\mathrm{cosh}(\beta x+\beta y)\right]}+_{1d}^{RS}[\pi ]+_{1d}^{RS}[\widehat{\pi }]`$ (4.8)
$`_{1d}^{RS}[\pi ]`$ $``$ $`\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{\beta N}}{\displaystyle ๐๐ซ[๐ฑ]๐\pi [๐]\mathrm{log}Z_{1d}^{RS}[๐ฑ,๐]}`$ (4.9)
$`Z_{1d}^{RS}[๐ฑ,๐]`$ $``$ $`{\displaystyle \underset{๐}{}}\mathrm{exp}\left[\beta {\displaystyle \underset{i=1}{\overset{N}{}}}(J_i+x_i)ฯต_i(๐)+\beta {\displaystyle \underset{i=1}{\overset{N}{}}}h_i\eta _i(๐)\right]`$ (4.10)
The distributions $`\pi `$ and $`\widehat{\pi }`$ are normalized ( $`๐x\pi (x)=๐y\widehat{\pi }(y)=1`$) and satisfy the saddle point equation below:
$`\pi (y)=\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _{\mathrm{}}^+\mathrm{}}๐\widehat{\pi }[๐]{\displaystyle ๐๐ซ[๐ฑ]\delta \left(y\frac{1}{\beta }\text{arctanh}\left(ฯต_i(๐)_{๐ฑ,๐}\right)+x_i\right)}`$ (4.11)
which is identical to the fixed point equation corresponding to Eq. (2.9), if we suppose the order parameters to be real at the saddle point.
Equation (4.11) is unpractical since it involves the unknown distributions $`\pi (x)`$ and $`\widehat{\pi }(x)`$ infinitely many times. However due to the short range structure of the hamiltonians defined in Eq. (1.2), it can be rewritten as a simple integral equation. Obviously the precise form of this equation depends upon the form of the hamiltonian (1.2). In particular it becomes simpler as the range of the interaction becomes shorter. Let us illustrate this point by considering the model (a) of the previous Section: $`ฯต_i(๐)=\sigma _i\sigma _{i1}`$, $`\eta _i(๐)=\sigma _i`$. We start by defining the following (right and left) partition functions:
$`Z_{i,M}^{(R)}(\sigma _i)`$ $``$ $`{\displaystyle \underset{\sigma _{i+1}\mathrm{}\sigma _{i+M}}{}}\mathrm{exp}\left[\beta {\displaystyle \underset{k=i+1}{\overset{i+M}{}}}(J_k+x_k)\sigma _k\sigma _{k1}+\beta {\displaystyle \underset{k=i}{\overset{i+M}{}}}h_k\sigma _k\right]`$ (4.12)
$`Z_{i,M}^{(L)}(\sigma _i)`$ $``$ $`{\displaystyle \underset{\sigma _{iM}\mathrm{}\sigma _{i1}}{}}\mathrm{exp}\left[\beta {\displaystyle \underset{k=iM+1}{\overset{i}{}}}(J_k+x_k)\sigma _k\sigma _{k1}+\beta {\displaystyle \underset{k=iM}{\overset{i}{}}}h_k\sigma _k\right]`$ (4.13)
and the (right and left) fields:
$`x_i^{R/L}\underset{M\mathrm{}}{lim}{\displaystyle \frac{1}{2\beta }}\mathrm{log}{\displaystyle \frac{Z_{i,M}^{(R/L)}(+)}{Z_{i,M}^{(R/L)}()}}`$ (4.14)
We define now a new couple of order parameters, the probability distributions $`\omega (x)`$ and $`\widehat{\omega }(x)`$ of the right (or left) fields:
$`\omega (x)={\displaystyle \underset{i0}{}dh_iP(h_i)\underset{i1}{}dJ_iP(J_i)\underset{i1}{}dx_i\pi (x_i)\delta \left(xx_0^R[J_i;h_i;x_i]\right)}`$ (4.15)
It easy to show that, at the saddle point, $`\omega (x)`$ and $`\widehat{\omega }(x)`$ satisfy the following integral equation:
$`\omega (z)`$ $`=`$ $`{\displaystyle ๐hP(h)๐J_1P(J_1)๐J_2P(J_2)๐x_1\widehat{\omega }(x_1)๐x_2\widehat{\omega }(x_2)๐z^{}\omega (z^{})}`$ (4.16)
$`\delta \left\{zh\mathrm{\Theta }_\beta [z^{};J_1+J_2+\mathrm{\Theta }_\beta (x_1;x_2)]\right\}`$
$`\mathrm{\Theta }_\beta (x;y)`$ $``$ $`{\displaystyle \frac{1}{\beta }}\text{arctanh}[\mathrm{tanh}(\beta x)\mathrm{tanh}(\beta y)]`$ (4.17)
and that the solution of Eq. (4.11) is related to the solution of the previous equation as follows:
$`\pi (x)={\displaystyle ๐JP(J)๐x_L\widehat{\omega }(x_L)๐x_R\widehat{\omega }(x_R)\delta [xJ\mathrm{\Theta }_\beta (x_R;x_L)]}`$ (4.18)
Equation (4.16) reduces to the Dyson Schmidt equation for a one dimensional model with nearest neighbour interaction if we keep the distribution $`\widehat{\omega }(x)`$ fixed. The interaction between the two one-dimensional subsystems turns it into a nonlinear equation. Moreover Eq. (4.16) can be treated numerically more easily than Eq. (4.11). A possible approach consists in representing the unknown distribution as $`\omega (x)=_{j=1}^K\delta (xx_j)`$ and iterating Eq. (4.16) until a fixed point is reached. An example of this kind of computations is shown in Fig. (2).
It is simple to obtain the analogous of Eq. (4.16) for the simplest non recursive model, defined by: $`ฯต_i(๐)=\sigma _i`$, $`\eta _i(๐)=\sigma _i\sigma _{i1}`$. The final result is
$`\omega (z)`$ $`=`$ $`{\displaystyle ๐hP(h)๐J_1P(J_1)๐J_2P(J_2)๐x_1\widehat{\omega }(x_1)๐x_2\widehat{\omega }(x_2)๐z^{}\omega (z^{})}`$ (4.19)
$`\delta \left(z\mathrm{\Theta }_\beta [h;J_1+J_2+x_1+x_2+z^{}]\right)`$
$`\pi (x)`$ $`=`$ $`{\displaystyle ๐JP(J)๐x_L\widehat{\omega }(x_L)๐x_R\widehat{\omega }(x_R)\delta (xJx_Lx_R)}`$ (4.20)
A simple approximation to the solution of Eq. (4.19) can be obtained by starting from a distribution $`\omega (x)`$ supported on very large fields $`x`$ and iterating Eq. (4.19) one time. The result is $`\pi (x)๐JP(J)๐h_1P(h_1)๐h_1P(h_1)\delta (xJh_1h_2)`$, which coincides with the more general Eq. (3.24) after the change of variables $`X=e^{2\beta x}`$. No such approximation is possible for Eq. (4.16).
Expressions equivalent to Eqs. (4.16-4.19) can be derived for more complicated types of interaction. In general the distribution $`\omega (x)`$, which is defined on the real line, will be replaced by a distribution defined on $`^{2^r1}`$, $`r`$ being the range of the hamiltonian.
## 5 The stability of the frozen solution.
We would like to study local stability of the no-error phase in the context of the replica method. This can be done<sup>6</sup><sup>6</sup>6For similar calculations see Ref. . by computing the eigenvalues of the matrices:
$`M_{a_1\mathrm{}a_l,b_1\mathrm{}b_m}^\pm [q]=\delta _{a_1\mathrm{}a_l,b_1\mathrm{}b_m}\pm {\displaystyle \frac{^2\beta _{1d,n}[q]}{q_{a_1\mathrm{}a_l}q_{b_1\mathrm{}b_m}}}`$ (5.1)
$`M^\pm [q]`$ are the mass matrices for purely real ($`M^+[q]`$), or purely immaginary ($`M^{}[q]`$), fluctuations of the order parameter around the value $`q`$. We are interested in the saddle point $`q_{a_1\mathrm{}a_l}^{}=2^{n/2}`$. In order to write down all the $`2^n`$ eigenvectors of $`M^\pm [q^{}]`$ it is convenient to change slightly our notation for the overlaps. Let us denote by $`\mathrm{\Omega }\{1,2,\mathrm{},n\}`$ the set of $`l|\mathrm{\Omega }|`$ different indices $`(a_1^\mathrm{\Omega },\mathrm{},a_l^\mathrm{\Omega })`$. We can use the $`\mathrm{\Omega }`$โs as indices for the overlaps with the natural identification $`q_\mathrm{\Omega }q_{a_1^\mathrm{\Omega }\mathrm{}a_l^\mathrm{\Omega }}`$. It is not difficult to show that
$`T_{\mathrm{\Omega }_a,\mathrm{\Omega }_b}^{(N)}`$ $``$ $`{\displaystyle \frac{1}{N}}{\displaystyle \frac{^2\mathrm{log}Z_{1d,n}[q]}{q_{a_1\mathrm{}a_l}q_{b_1\mathrm{}b_m}}}|_{q=q^{}}=`$
$`=`$ $`{\displaystyle \frac{2^{1n}}{N}}\left\{{\displaystyle \underset{(i,j)}{}}{\displaystyle ๐๐ซ[๐ฑ]e^{nN\beta fn\beta E_0}\left(1+e^{\beta \mathrm{\Delta }(i,j)}\right)^n\left[\mathrm{tanh}\left(\frac{\beta \mathrm{\Delta }(i,j)}{2}\right)\right]^u}{\displaystyle \frac{N^2}{2}}\right\}`$
where $`\mathrm{\Delta }(i,j)`$ is defined in Section 3, $`e^{nN\beta f}๐๐ซ[๐ฑ]e^{n\beta H(๐_0)}`$ and $`uu_{a_1\mathrm{}a_l,b_1\mathrm{}b_m}`$ counts the indices which are either in the set in $`\mathrm{\Omega }_a(a_1,\mathrm{},a_l)`$ or in the set $`\mathrm{\Omega }_b(b_1,\mathrm{},b_m)`$ but not in both. If $`q`$ is an eigenvector of $`T^{(N)}`$ with eigenvalue $`\theta _N`$, then it is an eigenvector of $`M^\pm [q^{}]`$ with eigenvalue $`\mu ^\pm =1lim_N\mathrm{}\theta _N`$.
Notice that $`T^{(N)}`$ is an hermitian matrix with respect to to the scalar product:
$`q,q^{}_n{\displaystyle \underset{l=0}{\overset{n}{}}}{\displaystyle \underset{(a_1,\mathrm{},a_l)}{}}q_{a_1\mathrm{}a_l}^{}q_{a_1\mathrm{}a_l}^{}={\displaystyle \underset{\mathrm{\Omega }}{}}q_\mathrm{\Omega }^{}q_\mathrm{\Omega }^{}`$ (5.3)
We shall use another subset of $`\{1,\mathrm{},n\}`$ (let us call it $`\mathrm{\Lambda }`$) to label the different eigenvectors of $`T^{(N)}`$, which we now exhibit:
$`q_\mathrm{\Omega }^{(\mathrm{\Lambda })}{\displaystyle \frac{1}{2^{n/2}}}(1)^{|\mathrm{\Lambda }\mathrm{\Omega }|}`$ (5.4)
The vectors $`\{q^{(\mathrm{\Lambda })}\}`$ form an orthonormal set with respect to the scalar product defined in Eq. (5.3). This is easily proven by induction on $`n`$. The vector $`q^{(\mathrm{})}`$ is nothing but the constant one. The corresponding eigenvalue is $`\theta _N^{(\mathrm{})}=1`$, whence $`\mu _{(\mathrm{})}^+=2`$ and $`\mu _{(\mathrm{})}^{}=0`$. The eigenvalue $`\mu _{(\mathrm{})}^{}=0`$ is a remnant of the invariance of the action under the symmetry cited in the footnote 5 of the previous Section. In order to compute the eigenvalues in the subspace orthogonal to $`q^{(\mathrm{})}`$, the following formula turns out to be useful:
$`{\displaystyle \underset{\mathrm{\Omega }^{}}{}}x^{|\mathrm{\Omega }\mathrm{\Omega }^{}|}q_\mathrm{\Omega }^{}^{(\mathrm{\Lambda })}=(1x)^{|\mathrm{\Lambda }|}(1+x)^{n|\mathrm{\Lambda }|}q_\mathrm{\Omega }^{(\mathrm{\Lambda })}`$ (5.5)
where $`\mathrm{\Omega }\mathrm{\Omega }^{}`$ denotes the symmetric difference of $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ (i.e. $`\mathrm{\Omega }\mathrm{\Omega }^{}(\mathrm{\Omega }\mathrm{\Omega }^{})(\mathrm{\Omega }^{}\mathrm{\Omega })`$). Using Eq. (5.5) and the results of algebra outlined in Section 3 we get (for $`\mathrm{\Lambda }\mathrm{}`$):
$`\theta _N\mathrm{}(\mathrm{\Lambda })=2\zeta _J^2{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\zeta _h^{\mathrm{weight}(s_m)}`$ (5.6)
where $`\mathrm{weight}(s_m)`$ is defined in Section 3 and
$`\zeta _C=\zeta _C(|\mathrm{\Lambda }|,n,\beta )={\displaystyle \frac{{\displaystyle ๐CP(C)e^{(n2|\mathrm{\Lambda }|)\beta C}}}{{\displaystyle ๐CP(C)e^{n\beta C}}}}`$ (5.7)
for $`Ch`$ or $`CJ`$. When the one dimensional hamiltonians (1.2) satisfy the hypothesis of Lemma 3.3, the sum in Eq. (5.6) can be explicitly computed yielding:
$`\theta _N\mathrm{}(\mathrm{\Lambda })={\displaystyle \frac{2\zeta _J^2\zeta _h^{w_0(f_1,f_2)+w_1(f_1,f_2)}}{1\zeta _h^{w_1(f_1,f_2)}}}`$ (5.8)
If $`n2|\mathrm{\Lambda }|`$ then $`\theta (\mathrm{\Lambda },n;\beta )`$ is positive and decreasing with $`\beta `$. Moreover $`lim_\beta \mathrm{}\theta (\mathrm{\Lambda },n;\beta )=0`$ and $`lim_{\beta 0}\theta (\mathrm{\Lambda },n;\beta )=\mathrm{}`$. We can thus define the critical temperatures $`\beta _{l,n}`$ for $`n/2l=|\mathrm{\Lambda }|1`$, by requiring <sup>7</sup><sup>7</sup>7Notice that Eq. (5.9) can have more than one solution for $`n<2|\mathrm{\Lambda }|`$. The โphysicalโ critical point is obtained by taking the limit $`n0`$ of the solution of Eq. (5.9) which exists for any $`n`$.
$`\theta (\mathrm{\Lambda },n;\beta _{|\mathrm{\Lambda }|,n})=1`$ (5.9)
If $`\beta >\beta _{|\mathrm{\Lambda }|,n}`$ the โfrozenโ saddle point is stable with respect to the direction $`q^{(\mathrm{\Lambda })}`$. If $`\beta <\beta _{|\mathrm{\Lambda }|,n}`$ it becomes unstable: $`\mu _{(\mathrm{\Lambda })}^+=1\theta (\mathrm{\Lambda })<0`$ while $`\mu _{(\mathrm{\Lambda })}^{}=1+\theta (\mathrm{\Lambda })>0`$ (it could be guessed that the โimaginaryโ directions would be stable because of the physical interpretation of the overlaps). In the limit $`n0`$, $`\beta _{l,n}\beta _c/l`$: the critical directions are the ones corresponding to $`|\mathrm{\Lambda }|=1`$. It is easy to see that the critical temperature $`\beta _c`$ coincides with the one obtained in Section 3.
## 6 Conclusion.
We have presented two derivations of the local stability condition for the no-error phase. Both will be object of the criticism of the skeptical reader. In the first one we obtained the โmean fieldโ equation describing the dynamics of the decoding algorithm, Eq. (2.9), by making use of heuristic arguments. Indeed we argued Eq. (2.9) to be valid only in the replica symmetric approximation. In the second derivation we made use of the replica method, which has not (yet) well founded mathematical basis.
We think that the two derivations compensate each other for their defects. Moreover they yield the same replica symmetric saddle point equation (4.11) and give the same picture of the instability which destroys the no-error (frozen) phase. This corresponds to couples of flipped $`ฯต(๐)`$โs. Finally thanks to the first derivation we get some insight on the behavior of the decoding algorithm. In particular we have seen that, in the frozen phase, it approaches a no-error fixed point. This approach becomes slower near to the boundary of the frozen phase.
In Ref. the local stability threshold computed here has been compared with numerical simulations for two types of code, respectively the models (a) and (b) presented in Section 3. Good agreement was found only for model (a). We propose two possible explanations of the disagreement for model (b):
* the phase transition is a first order one;
* the turbo decoding algorithm used in Ref. gets sticked in some local minimum of the free energy, characterized by a finite error probability per bit.
We have not yet enough informations for choosing between these two scenarios.
## Appendix A Useful algebra results.
In this Appendix we remind to the reader some known facts in the theory of finite fields and we prove the propositions stated in Section 3. These are nothing but simple exercises and we work out them in detail only for greater convenience of the reader. Finally we illustrate a few applications of the results obtained. The reader interested in a more complete treatment can consult Refs. .
Let us begin with some elementary definitions. The basic object is $`_2`$ i.e. the field of integer numbers modulo $`2`$. A polynomial over $`_2`$, $`f(x)_2[x]`$ is simply a polynomial whose coefficients are in $`_2`$. We say $`f(x)_2[x]`$ to be irreducible if there do not exist two noncostant polynomials $`g(x),h(x)_2[x]`$ such that $`f(x)=g(x)h(x)`$. Any $`f(x)_2[x]`$ possess an unique factorization, i.e. a decomposition of the form $`f(x)=f_1(x)^{r_1}\mathrm{}f_h(x)^{r_h}`$ where $`f_i(x)_2[x]`$ are irreducible and $`r_i1`$ are integer numbers. Given two polynomials $`f(x),g(x)_2[x]`$ we say that $`f(x)`$ divides $`g(x)`$ (in symbols $`f(x)|g(x)`$) if there exists $`h(x)_2[x]`$ such that $`g(x)=f(x)h(x)`$ <sup>8</sup><sup>8</sup>8Similarly, given two integer numbers $`p,q`$, we say that $`p`$ divides $`q`$ (and write $`p|q`$) if there exists $`m`$, such that $`q=mp`$.. For an irreducible polynomial $`f(x)_2[x]`$ it does make sense to define the order $`o(f)`$: $`o(f)`$ is the smallest positive integer $`k`$ such that $`f(x)|x^k+1`$. The basic result which we shall employ in this Appendix is the following:
###### Theorem A.1
Let $`f(x)`$ be an irreducible polynomial over $`_2`$. Then $`f(x)|x^k+1`$ if and only if $`o(f)|k`$.
It is useful to know how to compute the order of an irreducible polynomial. The main tool is the theorem below:
###### Theorem A.2
Let $`f(x)`$ be an irreducible polynomial of degree $`d`$ over $`_2`$. Then $`d`$ is the smallest positive integer for which $`o(f)|2^d1`$.
Moreover it is obvious from the definition that $`o(f)\mathrm{deg}(f)`$
Our first step will be the proof of Lemma 3.2 which we restate here as follows
###### Lemma A.1
Let $`f(x)`$ be a polynomial on $`_2`$ with the following factorization
$`f(x)=f_1^{r_1}(x)\mathrm{}f_h^{r_h}(x);r_i1`$ (A.1)
where the polynomials $`f_i(x)`$ are irreducible over $`_2`$. Let $`p_i`$ be the smallest integer such that $`2^{p_i}r_i`$. Then $`f(x)|(1+x^k)`$ if and only if $`2^{p_i}|k`$ and $`o(f_i)|k`$ for $`i\{1,\mathrm{},h\}`$.
Proof of Lemma A.1. Let us begin by noticing that, since the $`f_i(x)`$ are irreducible, $`f(x)|(1+x^k)`$ if and only if $`f_i^{r_i}(x)|(1+x^k)`$ for $`i\{1,\mathrm{},h\}`$. We can then limit ourselves to the case $`f(x)=h^r(x)`$ with $`h(x)`$ irreducible. It is convenient to work in an extension of $`_2`$, i.e. in a field containing $`_2`$ as a subfield. We choose an extension (let us call it $`S`$) of $`_2`$ such that both $`h(x)`$ and $`(1+x^k)`$ can be decomposed in linear factors. The existence of such an extension is a basic fact of field theory. We are then looking for the $`k`$ such that all the root of $`h(x)`$ (in $`S`$) are roots of $`(1+x^k)`$ with multiplicity at least $`r`$. It is then necessary to study the multiplicity of the roots of $`(1+x^k)`$. The first observation is that, if $`k`$ is odd, all the roots are simple. In fact $`\frac{d}{dx}(1+x^k)=kx^{k1}`$ has no roots in common with $`(1+x^k)`$. The second observation consists in noticing that $`(1+x^{2k})=(1+x^k)^2`$. We deduce that $`(1+x^{2^mk})`$ with $`k`$ odd has $`k`$ distinct roots (the same as $`(1+x^k)`$), each one with multiplicity $`2^m`$. The final outcome is that $`h^r(x)|(1+x^{2^mk})`$ if and only if $`2^mr`$ and $`o(h)|k`$ $`\mathrm{}`$
From Lemma A.1 the explicit form of the period $`\omega (f)`$ used in Sec. 3 is easily obtained:
$`\omega (f)=2^{\mathrm{max}(p_1,\mathrm{},p_h)}\mathrm{lcm}(o(f_1),\mathrm{},o(f_h))`$ (A.2)
The Lemmas 3.1 and 3.3 are easy consequences of Lemma 3.2.
Proof of Lemma 3.1. Let us begin by considering the series $`1/f(x)`$. We can always define the polynomials $`\phi _n(x)`$ with $`\mathrm{deg}(\phi _n)<\omega (f)`$ such that $`1/f(x)=_{n=0}^{\mathrm{}}\phi _n(x)x^{n\omega (f)}`$. Since $`f(x)`$ divides $`(1+x^{m\omega (f)})`$ for all $`m1`$, we conclude that $`\phi _n(x)=\phi _n^{}(x)\phi (x)`$ for all $`n,n^{}0`$ and $`(1+x^{m\omega (f)})/f(x)=_{k=0}^{m1}\phi (x)x^{k\omega (f)}`$. With the following definition
$`g(x)\phi (x){\displaystyle \underset{l=0}{\overset{L}{}}}g_l(x)x^{l\omega (f)},\mathrm{deg}[g_l(x)]<\omega (f)`$ (A.3)
we get
$`{\displaystyle \frac{g(x)}{f(x)}}={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}x^{n\omega (f)}{\displaystyle \underset{l=0}{\overset{\mathrm{min}(n,L)}{}}}g_l(x){\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}x^{n\omega (f)}p_n(x)`$ (A.4)
Notice that, for $`nL`$, $`p_n(x)=p_{\mathrm{}}(x)g(x)/f(x)modx^{\omega (f)}`$. An upper bound on $`L`$ is easily obtained from Eq. (A.3) yielding<sup>9</sup><sup>9</sup>9Here use the definition $`x\mathrm{min}\{n:n>x\}`$. $`p_n(x)=p_{\mathrm{}}(x)`$ for $`n(\mathrm{deg}[g(x)]1)/\omega (f)L`$. Clearly it cannot be $`p_{\mathrm{}}(x)=0`$ otherwise we would conclude that $`f(x)`$ divides $`g(x)`$ in contradiction with the hypothesis. In order to complete the proof of let us suppose the following equation to hold
$`{\displaystyle \frac{g(x)}{f(x)}}={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}x^{n\omega ^{}}p_n^{}(x)`$ (A.5)
with $`\omega ^{}<\omega (f)`$, $`\mathrm{deg}(p_n^{})<\omega ^{}`$ and $`p_n^{}(x)=p_{\mathrm{}}^{}(x)`$ for $`n`$ large enough. This implies that $`f(x)`$ divides $`g(x)(1+x^\omega ^{})`$ but, since $`\mathrm{gcd}(f,g)=1`$, we would conclude that $`f(x)`$ divides $`(1+x^\omega ^{})`$ contradicting Lemma 3.2 $`\mathrm{}`$
Proof of Lemma 3.3. It suffices to specialize the content of the previous paragraph to the case $`\mathrm{deg}[g(x)]\omega (f)`$:
$`g(x)\phi (x)`$ $`=`$ $`g_0(x)+g_1(x)x^{\omega (f)}`$ (A.6)
$`s_m(x){\displaystyle \frac{g(x)}{f(x)}}(1+x^{m\omega (f)})`$ $`=`$ $`g_0(x)+\{g_0(x)+g_1(x)\}{\displaystyle \underset{h=1}{\overset{m1}{}}}x^{h\omega (f)}+g_1(x)`$ (A.7)
whence
$`\text{weight}[s_m(x)]`$ $`=`$ $`w_0+w_1m`$ (A.8)
$`w_0`$ $``$ $`\text{weight}[g_0(x)]+\text{weight}[g_1(x)]\text{weight}[g_0(x)+g_1(x)]`$ (A.9)
$`w_1`$ $``$ $`\text{weight}[g_0(x)+g_1(x)]`$ (A.10)
$`\mathrm{}`$
What does it happen when the hypothesis of Lemma 3.3 are not satisfied? It is easy to guess the answer. There exists a positive integer $`m_0`$ such that, for $`mm_0`$, $`\mathrm{weight}(s_m)`$ grows linearly with $`m`$: $`\mathrm{weight}(s_m)=\stackrel{~}{w}_0(f,g)+\stackrel{~}{w}_1(f,g)m`$ with $`\stackrel{~}{w}_1(f,g)=\mathrm{weight}(p_{\mathrm{}})`$. Thanks to this fact we can always sum the series in Eq. (3) in the interval $`0<\beta <\beta _1`$. The discussion of the behavior of Eq. (3.8) presented in Section 3 is then completely general.
Let us return down to the earth and make a few examples. We shall consider the codes presented in Ref. :
1. The simplest non trivial case: $`f(x)=1+x`$, $`g(x)=1`$. Clearly both the polynomials are irreducible. The degree of $`f(x)`$ is $`\mathrm{deg}[f(x)]=1`$. Because of Theorem A.2 $`o(f)|2^11=1`$ whence $`o(f)=1=\omega (f)`$. Theorem A.1 implies that $`f(x)|1+x^k`$ for each $`k1`$. This conclusion is easily confirmed by the well known formula $`(1+x^k)=(1+x)(1+x+\mathrm{}+x^{k1})`$. Lemma 3.1 tells us that $`g(x)/f(x)=_{n=0}^{\mathrm{}}p_nx^n`$ with $`p_n=p_{\mathrm{}}`$ for $`n0`$ and that $`p_{\mathrm{}}=1`$ ($`1`$ is the unique non zero polynomial of degree zero). We have thus rediscovered the simple fact that $`(1+x)^1=_{n=0}^{\mathrm{}}x^n`$. Finally we observe the hypothesis of Lemma 3.3 are satisfied and that (with the notation of Eq. (A.6)), $`g_0(x)=1`$ and $`g_1(x)=0`$. From Eqs. (A.8-A.9-A.10) it follows that $`\text{weight}[s_m(x)=(1+x^m)/(1+x)]=m`$ which is easily confirmed by observing that $`s_m(x)=1+x+\mathrm{}+x^{m1}`$.
2. A less elementary example is: $`f(x)=1+x+x^2`$, $`g(x)=1+x^2`$. It is easy to see that $`f(x)`$ is irreducible and that $`g(x)=(1+x)^2`$ whence $`\mathrm{gcd}(f,g)=1`$. From $`o(f)|2^{\mathrm{deg}(f)}1=3`$ and $`o(f)\mathrm{deg}(f)=2`$ we deduce that $`o(f)=3=\omega (f)`$. In fact
$`{\displaystyle \frac{1}{1+x+x^2}}`$ $`=`$ $`1+x+x^3+x^4+x^6+\mathrm{}={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\phi (x)x^{3n}`$ (A.11)
$`\phi (x)`$ $`=`$ $`1+x`$ (A.12)
Thus by Lemma 3.2 $`f(x)|(1+x^k)`$ if and only if $`k`$ is a multiple of $`3`$. We can use Lemma 3.3 in order to compute the weight of $`s_m(x)=(1+x^2)(1+x^{3m})/(1+x+x^2)`$. We see that $`g_0(x)=1+x+x^2`$ and $`g_1(x)=1`$ whence $`\text{weight}[h_m(x)]=2+2m`$. With some book-keeping one can confirm this result:
$`h_m(x)=x+{\displaystyle \underset{l=0}{\overset{m1}{}}}(x^{3l+1}+x^{3l+2})+x^{3m}`$ $``$ $`\text{weight}[h_m(x)]=2+2m`$ (A.13)
3. Finally the generating polynomials used in Ref. to build the first example of turbo code: $`f(x)=1+x+x^2+x^3+x^4`$, $`g(x)=1+x^4`$. Also in this case $`f(x)`$ is irreducible and $`g(x)=(1+x)^4`$ yielding $`\mathrm{gcd}(f,g)=1`$. Since $`o(f)|2^{\mathrm{deg}(f)}1=15`$ and $`o(f)\mathrm{deg}(f)=4`$, we deduce that either $`o(f)=5`$ or $`o(f)=15`$. However we know that $`(1+x^5)=(1+x)(1+x+x^2+x^3+x^4)`$ and we conclude that $`o(f)=5=\omega (f)`$. In fact
$`{\displaystyle \frac{1}{1+x+x^2}}`$ $`=`$ $`1+x+x^5+x^6+x^{10}+x^{11}+\mathrm{}={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\phi (x)x^{5n}`$ (A.14)
$`\phi (x)`$ $`=`$ $`1+x`$ (A.15)
Using the fact that $`g_0(x)=1+x+x^4`$ and $`g_1(x)=1`$ we get $`\text{weight}[h_m(x)]=2m+2`$.
## Appendix B On the asymptotic behavior of the solutions of Eq. (3.8).
In this Appendix we study Eq. (3.8) in order to extend to this case the results concerning Eq. (3.9) used in Section 3. We shall examine both the approach of Ref. , which is based on the analogy with the KPP equation and is a non rigorous one, and the approach of Ref. , which employs probability theory and is entirely satisfactory from the mathematical point of view.
We would like to deal with this type of equation:
$`Q_{n+1}(Z)`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}P(V)๐V{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}p(h)dh{\displaystyle _0^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}Q_n(Z_i)dZ_i`$ (B.1)
$`\delta \left(Z{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\mathrm{exp}\left\{\beta V\beta {\displaystyle \underset{j=1}{\overset{i1}{}}}h_j\right\}Z_i\right)`$
with the requirement that $`๐hp(h)h>0`$ and the initial condition $`P_0(Z)=\delta (Z1)`$. Following Ref. we introduce the function:
$`G_n(x){\displaystyle _0^{\mathrm{}}}๐ZQ_n(Z)\mathrm{exp}\{e^{\beta x}Z\}`$ (B.2)
which satisfy this recurrence equation
$`G_{n+1}(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}P(V)๐V{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}p(h_i)dh_i{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}G_n(x+V+{\displaystyle \underset{j=1}{\overset{i1}{}}}h_j)`$ (B.3)
Let us make a few elementary observations concerning Eq. (B.3): if $`0G_m(x)1`$ for some $`m`$ and all $`x`$ then $`0G_n(x)1`$ for all $`x`$ and $`n>m`$; if $`limsup_x\mathrm{}G_n(x)=g_{\mathrm{}}<1`$ then $`G_{n+1}=0`$; if $`G_n(x)`$ is increasing and $`0<G_n(x)<1`$ for some $`x`$ (both these hypothesis are implied by Eq. (B.2)) then $`lim_x\mathrm{}G_{n+1}(x)=0`$. The stationary uniform solutions of Eq. (B.3) are $`G_n^A(x)=0`$ and $`G_n^B(x)=1`$. The first one is obviously stable. If we consider a small fluctuation around $`G_n^B(x)`$, $`G_n(x)=1+\rho _n(x)`$ we get:
$`\rho _{n+1}(x){\displaystyle _{\mathrm{}}^{\mathrm{}}}P(V)๐V{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}p(h)dh{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\rho _n(x+V+{\displaystyle \underset{j=1}{\overset{i1}{}}}h_j)`$ (B.4)
which can be diagonalized in Fourier space:
$`\widehat{\rho }_{n+1}(k){\displaystyle \frac{\widehat{P}(k)}{1\widehat{p}(k)}}\widehat{\rho }_n(k)\lambda (k)\widehat{\rho }_n(k)`$ (B.5)
Notice that $`|\lambda (k)|>1`$ for $`k`$ small enough and that $`|\lambda (k)|`$ diverges at $`k=0`$ in agreement with the previous observation that if $`G_n(x)=1\rho `$ than $`G_{n+1}(x)=0`$. The preceding observations lead us to the hypothesis that the $`n\mathrm{}`$ behavior of our problem is controlled by front-like solutions $`G_n(x)=g(xc(\beta )n)`$ interpolating between the stable state $`G_n^A(x)=0`$ at $`x\mathrm{}`$ and $`G_n^B(x)=1`$ at $`x\mathrm{}`$.
This scenario is easily confirmed in the case without disorder. If $`P(V)=\delta (VV_0)`$ and $`p(h)=\delta (hh_0)`$ one obtains $`P_n(Z)=\delta (Ze^{\beta c(\beta )n})`$, $`G_n(x)=\mathrm{exp}\{e^{\beta (xc(\beta )n)}\}`$ with
$`c(\beta )={\displaystyle \frac{1}{\beta }}\mathrm{log}{\displaystyle \frac{e^{\beta V_0}}{1e^{\beta h_0}}}`$ (B.6)
In the general case we assume the existence of front-like solutions with the large $`x`$ behavior $`G_n(x)1e^{\beta (xc(\beta )n)}+o(e^{\beta x})`$. The front velocity is obtained through the construction given in Eqs. (3.14-3.15) with
$`v(\beta ){\displaystyle \frac{1}{\beta }}\mathrm{log}\varphi (\beta )={\displaystyle \frac{1}{\beta }}\mathrm{log}{\displaystyle \frac{e^{\beta V}}{1e^{\beta h}}}`$ (B.7)
Notice that $`h>0`$ implies that $`e^{\beta h}<1`$ in some interval $`0<\beta <\beta _1`$ and that $`\beta _c<\beta _1`$. This remark allows us to sum the series $`_ke^{\beta h}^k`$ in the range $`0<\beta <\beta _c`$, thus obtaining Eq. (B.7). The same remark will be useful in the following.
Let us consider now the more rigorous approach used in Ref. . We start by defining the polymer model which corresponds to Eq. (B.1). We have to use a tree with a numerable set of branches stemming from each node. A node of the $`n`$-th generation is identified by $`n`$ integer numbers $`\underset{ยฏ}{\omega }(\omega _1,\mathrm{},\omega _n)`$; its generation is denoted by $`|\underset{ยฏ}{\omega }|`$. We denote by $`\underset{ยฏ}{0}`$ the root node (i.e. the only node of the zeroth generation). We say that the node $`\underset{ยฏ}{\omega }^{}`$ belonging to the $`m`$-th generation is a descendant of the node $`\underset{ยฏ}{\omega }`$ of the $`n`$-th generation (and write $`\underset{ยฏ}{\omega }\underset{ยฏ}{\omega }^{}`$ if $`n<m`$ or $`\underset{ยฏ}{\omega }\underset{ยฏ}{\omega }^{}`$ if $`nm`$ ) if $`\omega _1=\omega _1^{},\mathrm{},\omega _n=\omega _n^{}`$. The node $`\underset{ยฏ}{\omega }^{}`$ is said to be an older brother of the node $`\underset{ยฏ}{\omega }`$ with $`|\underset{ยฏ}{\omega }^{}|=|\underset{ยฏ}{\omega }|=n`$ if $`\omega _1=\omega _1^{},\mathrm{},\omega _{n1}=\omega _{n1}^{}`$ and $`\omega _n>\omega _n^{}`$. A pair of random variables $`V(\underset{ยฏ}{\omega })`$ and $`h(\underset{ยฏ}{\omega })`$ is attached at each node. All these variables are statistically independent and have marginal distributions $`p(h)`$ (the $`h(\underset{ยฏ}{\omega })`$โs) and $`P(V)`$ (the $`V(\underset{ยฏ}{\omega })`$โs). A directed polymer is given by a pair of nodes $`\underset{ยฏ}{\omega }^1\underset{ยฏ}{\omega }^2`$. To each polymer we assign an energy as follows:
$`E(\underset{ยฏ}{\omega }^1,\underset{ยฏ}{\omega }^2)={\displaystyle \underset{\underset{ยฏ}{\omega }^1\underset{ยฏ}{\omega }\underset{ยฏ}{\omega }^2}{}}V(\underset{ยฏ}{\omega })+{\displaystyle \underset{\underset{ยฏ}{\omega }^1\underset{ยฏ}{\omega }\underset{ยฏ}{\omega }^2}{}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^{}:\underset{ยฏ}{\omega }^{}\text{ is an older}}{\text{brother of }\underset{ยฏ}{\omega }}}{}}h(\underset{ยฏ}{\omega }^{})`$ (B.8)
Moreover we use the shorthand $`E(\underset{ยฏ}{0},\underset{ยฏ}{\omega })E(\underset{ยฏ}{\omega })`$ and define the following partition functions:
$`Z_n(\beta ){\displaystyle \underset{\underset{ยฏ}{\omega }:|\underset{ยฏ}{\omega }|=n}{}}e^{\beta E(\underset{ยฏ}{\omega })}`$ (B.9)
$`Z_n(\beta |\underset{ยฏ}{\omega })={\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^{}\underset{ยฏ}{\omega }:}{|\underset{ยฏ}{\omega }^{}||\underset{ยฏ}{\omega }|=n}}{}}e^{\beta E(\underset{ยฏ}{\omega },\underset{ยฏ}{\omega }^{})}`$ (B.10)
The velocity of the front wave studied in the previous paragraphs corresponds in this language to the random variable:
$`c(\beta )\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{n\beta }}\mathrm{log}Z_n(\beta )`$ (B.11)
The model has two phases. In the high temperature phase ($`\beta \beta _c`$) the fluctuations of $`Z_n(\beta )`$ are small and
$`c(\beta )=\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{n\beta }}\mathrm{log}Z_n(\beta )=v(\beta )`$ (B.12)
In the low temperature phase ($`\beta >\beta _c`$) the fluctuations become large and $`c(\beta )`$ is fixed by simple convexity and monotonicity arguments. The key point of the approach used in Ref. is to estimate these fluctuations by proving that, for $`\beta <\beta _c`$:
$`{\displaystyle \frac{Z_n(\beta )^\alpha }{Z_n(\beta )^\alpha }}\text{Bound}(\alpha ,\beta )`$ (B.13)
for some $`1<\alpha <2`$ uniformly in $`n`$. This is enough for obtaining Eq. (B.12).
Let us define the normalized variables $`M_n(\beta )Z_n(\beta )/Z_n(\beta )`$. In Ref. the bound in Eq. (B.13) is obtained starting with the second moment of $`M_n(\beta )`$, and then refining the inequality for the fractional moments of order $`1<\alpha <2`$. Notice that looking at the $`m`$-th moment of the partition function is a well known method for obtaining an upper estimate on the critical temperature (the estimate becomes worser as $`m`$ gets larger). Let us have a look at the first two integer moments:
$`Z_{n+1}(\beta )`$ $`=`$ $`e^{\beta V}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}e^{\beta h}^kZ_n(\beta )`$ (B.14)
$`Z_{n+1}^2(\beta )`$ $`=`$ $`\left(e^{\beta V}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}e^{\beta h}^k\right)^2\left[Z_n^2(\beta )Z_n(2\beta )\right]+`$ (B.15)
$`+e^{2\beta V}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}e^{2\beta h}^k\left(1+2{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}e^{\beta h}^l\right)Z_n(2\beta )`$
In general the $`m`$-th moment is finite (but not necessarily uniformly bounded) only if $`e^{m\beta h}<1`$ i.e. if $`\beta <\beta _1/m`$. There is no integer moment of order greater than one which remains finite in the interval $`(0,\beta _c)`$. This fact forces us to a slight modification of the proof presented in Ref. . We use the trivial identity:
$`Z_{n+1}(\beta )={\displaystyle \underset{\underset{ยฏ}{\omega }:|\underset{ยฏ}{\omega }|=1}{}}e^{\beta E(\underset{ยฏ}{\omega })}Z_n(\beta |\underset{ยฏ}{\omega })`$ (B.16)
and estimate the $`\alpha `$-th moment (with $`1<\alpha <2`$) as follows:
$`Z_{n+1}^\alpha (\beta )`$ $`=`$ $`\left\{{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^1:}{|\underset{ยฏ}{\omega }^1|=1}}{}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^2:}{|\underset{ยฏ}{\omega }^2|=1}}{}}e^{\beta [E(\underset{ยฏ}{\omega }^1)+E(\underset{ยฏ}{\omega }^2)]}Z_n(\beta |\underset{ยฏ}{\omega }^1)Z_n(\beta |\underset{ยฏ}{\omega }^2)\right\}^{\alpha /2}`$ (B.17)
$``$ $`{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^1:}{|\underset{ยฏ}{\omega }^1|=1}}{}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^2:}{|\underset{ยฏ}{\omega }^2|=1}}{}}e^{\frac{\alpha \beta }{2}[E(\underset{ยฏ}{\omega }^1)+E(\underset{ยฏ}{\omega }^2)]}Z_n^{\alpha /2}(\beta |\underset{ยฏ}{\omega }^1)Z_n^{\alpha /2}(\beta |\underset{ยฏ}{\omega }^2)`$
For a temperature such that $`\alpha \beta <\beta _1`$ we can take the averages and sum up the series:
$`Z_{n+1}^\alpha (\beta )`$ $``$ $`{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }:}{|\underset{ยฏ}{\omega }|=1}}{}}e^{\alpha \beta E(\underset{ยฏ}{\omega })}Z_n^\alpha (\beta )+{\displaystyle \underset{\genfrac{}{}{0pt}{}{\underset{ยฏ}{\omega }^1\underset{ยฏ}{\omega }^2:}{|\underset{ยฏ}{\omega }^i|=1}}{}}e^{\frac{\alpha \beta }{2}[E(\underset{ยฏ}{\omega }^1)+E(\underset{ยฏ}{\omega }^2)]}Z_n^{\alpha /2}(\beta )^2`$ (B.18)
$``$ $`\varphi (\alpha \beta )Z_n^\alpha (\beta )+2\varphi (\alpha \beta ){\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}e^{\frac{\alpha \beta }{2}h}^lZ_n(\beta )^\alpha `$
Rewriting this formula for the normalized variables we get
$`M_{n+1}^\alpha (\beta )`$ $``$ $`\left[{\displaystyle \frac{\varphi (\alpha \beta )}{\varphi (\beta )^\alpha }}\right]M_n^\alpha (\beta )+2\left[{\displaystyle \frac{\varphi (\alpha \beta )}{\varphi (\beta )^\alpha }}\right]{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}e^{\frac{\alpha \beta }{2}h}`$ (B.19)
$``$ $`\psi (\alpha ,\beta )M_n^\alpha (\beta )+\chi (\alpha ,\beta )`$
At this point we observe, following Ref. , that, if $`\frac{dv}{d\beta }(\beta )<0`$ (i.e. $`\beta <\beta _c`$) then we can choose $`\alpha >1`$ such that $`\psi (\alpha ,\beta )<1`$. The condition to be imposed on $`\alpha `$ for obtaining this inequality is $`\alpha <\beta _c/\beta `$ (notice that this inequality implies the previous one $`\alpha <\beta _1/\beta `$). The desired bound is obtained by using Gronwall lemma together with the fact that $`M_0^\alpha (\beta )=1`$:
$`M_n^\alpha (\beta )\psi ^n(\alpha ,\beta )+{\displaystyle \frac{1\psi ^n(\alpha ,\beta )}{1\psi (\alpha ,\beta )}}\chi (\alpha ,\beta )1+{\displaystyle \frac{1}{1\psi (\alpha ,\beta )}}\chi (\alpha ,\beta )`$ (B.20) |
warning/0003/quant-ph0003103.html | ar5iv | text | # Effectively classical quantum states for open systems
## Abstract
Notions of robust and โclassicalโ states for an open quantum system are introduced and discussed in the framework of the isometric-sweeping decomposition of trace class operators. Using the predictability sieve proposed by Zurek, โquasi-classicalโ states are defined. A number of examples illustrating how the โquasi-classicalโ states correspond to classical points in phase space connected with the measuring apparatus are presented.
1. Introduction
Quantum mechanics, whose basic laws were formulated in the twenties, still remains the most fundamental theory we know. Although, it was originally conceived as a theory of atoms, it has shown a wide range of applicability, making it more and more evident that the formalism describes some general properties of Nature. However, despite its successes, there is still no consensus about its interpretation with the main questions being centered around the quantum measurements. Clearly, the existence of classical quantities which would allow to express the measurement results and explain the classical appearance of the macroscopic world is a fundamental problem in this matter. The standard explanation which states that, for example, the center-of-mass motion of a macroscopic object should be described by a narrow wave packet, well localized in both position and momentum, is not satisfactory. It still remains unanswered why such objects are represented by narrow wave packets, while the superposition principle allows the emergence of non-classical states as well. Moreover, measurement-like processes would necessarily produce such non-classical states, as in the infamous example of Schrรถdingerโs cat. A superposition of being dead and alive should produce an entirely new state, in the same sense as the superposition of $`K`$ meson and its antiparticle does.
The main reason for this annoying situation seems to be based on the assumption that it is possible to isolate systems from their environment. When we drop it as unjustified and consider quantum systems as open ones we obtain a new perspective for the understanding of the emergence of classical properties within the framework of open systems theory. This is the basic objective of the program of decoherence proposed by Zurek and further developed in . For a recent review of the subject and a wide range of references up to 1996 see . Decoherence is a process of continuous measurement-like interaction between a system and its environment which results in limiting the validity of the superposition principle in the Hilbert space of the system. In other words, the environment destroys the vast majority of superpositions in short time, and, in the case of macroscopic objects, almost instantaneously. This leads to the appearance of environment-induced superselection rules, which precludes all but a particular subset of states from stable existence. On the other hand, it singles out a preferred set of states which behave in an effectively classical, predictable manner. Generalizing this notion the predictability sieve was introduced (see also ). It is a procedure, which systematically explores states of an open quantum system in order to arrange them and next put on a list, starting with the most predictable ones and ending with those, which are most affected by the environment. Clearly, the states being on the top of the list can be thought of as โclassicalโ or โquasi-classicalโ ones.
In order to study decoherence, the analysis of the evolution of the reduced density matrix obtained by tracing out the environment variables is the most convenient strategy. If the interaction is such that the reduced density matrix becomes approximately diagonal in a particular basis (in the simplest case), then it is said that an environment-induced superselection structure has emerged. Generally, the procedure of tracing out environment variables, being the composition of a unitary automorphism with a conditional expectation, leads to a complicated integro-differential equation for the reduced statistical operator. However, for a large class of interesting physical phenomena we can derive, using certain limiting procedures, an approximate Markovian master equation for the reduced density matrix . More recently, the derivation of the master equation for the reduced density matrix of a system coupled linearly to an ohmic, subohmic and supraohmic environment at arbitrary temperature has been obtained in .
Usually, when deriving the master equation for the reduced density matrices it is assumed that the quantum system interacts with the environment, which is another quantum system and hence is also described in terms of quantum mechanics. However, it should be pointed out here that such an assumption stemming from the thinking of quantum mechanics as a universal theory, can be replaced by a more general one. Sometimes, it is more useful and natural to treat the external degrees of freedom as a classical system described by a commutative algebra of functions. In this approach ( see for discrete classical systems and for continuous ones) the evolution equation for the classical part is modified by the expectation value of some quantum observable while, at the same time, the Schrรถdinger unitary dynamics for the quantum subsystem is replaced by a dynamical semigroup of completely positive maps. Therefore, when we allow the quantum system to interact with its environment, then, regardless of the nature of this interaction, the evolution becomes dissipative, given by the Markovian master equation.
The loss of quantum coherence in the Markovian regime was established in a number of open systems , giving a clear evidence of environment-induced superselection rules. In a recent paper a thorough mathematical analysis of the superselection structure associated to an environment-induced semigroup was presented. It was achieved by the use of the isometric-sweeping decomposition, which singles out a subspace of density matrices, on which the semigroup acts in a reversible, unitary way, and sweeps out the rest of statistical states. The purpose of this paper is to pursue that investigation with a particular emphasis on the analysis of the classicality of states. We put the notion of pointer states, previously introduced and discussed by Zurek and other authors, into a general framework, and examine their properties. It is worth noting that proposed definitions are expressed solely in terms of the dynamical semigroup and does not refer to any additional conditions like that one involving the knowledge of the state of the quantum system before the beginning of interaction. A number of examples including that of a quantum stochastic process of Davies, and illustrating how the โquasi-classicalโ states correspond to classical points in the underlying phase space are also presented.
2. โClassicalโ states
One can show that purely unitary evolution can never resolve the apparent conflict between predictions it implies and perception of the classical reality. Therefore, in order to explain the appearance of classical (non-quantum) properties of a quantum system, we have to open the system and allow it to interact with the environment. As was mentioned in Introduction we restrict our considerations to the Markovian regime and thus assume that the evolution of the reduced density matrix is given by an environment-induced semigroup. The concept of the environment-induced semigroup was introduced in , and the justification of the name was also given there. They form a subclass of dynamical semigroups (completely positive, trace preserving and contractive in the trace norm $`_1`$), which are also contractive in the operator norm $`_{\mathrm{}}`$. This additional property ensures that both the linear and statistical entropy of the open quantum system never decrease in the course of interaction .
In this section we search for quantum states which can correspond to the classical points in phase space once the interaction with the environment is acknowledged. Our strategy is as follows. We start with a quantum system whose evolution is given by an environment-induced semigroup $`T_t`$ without asking question where it comes from. Then, using general principles and properties of that semigroup we determine sets of โclassicalโ and โquasi-classicalโ (see the next section) states. If they are empty and all states are stable, the system can be thought of as closed, evolving in a unitary way, and hence the semigroup may be extended to a one parameter group of unitary automorphisms. However, if one of them is non-trivial, then the system is open and we conclude that the selected states correspond with points in a classical phase space. Therefore, these sets contain the information about the type of interaction with the environment (measuring apparatus) which led to the appearance of the semigroup $`T_t`$.
At first, let us comment on crucial differences between states of classical and quantum systems. One of the most characteristic features distinguishing classical from quantum states is their sensitivity to measurements. In classical physics we could perform many kinds of measurements which would not disturb the system in an essential way. A measurement can increase our knowledge of the state of the system but, in principle, it has no effect on the system itself. By contrast, in quantum mechanics it is impossible to find out what the state is without, at the same time, changing it in the way determined by the measurement. According to the von Neumann projection postulate the outcome will be, in general, represented by a density matrix. Therefore, as a convenient measure of the influence of the environment on the state, we take the measure of the loss of its purity expressed in terms of the linear entropy $`S_{\mathrm{lin}}(\rho )=\mathrm{tr}(\rho \rho ^2)`$ .
Let $`๐ฎ`$ denote the set of all states of the quantum system. By a state we always mean a pure state, whereas for a mixed state we reserve such notions like density matrix or statistical state. Hence $`๐ฎ`$ consists of unit vectors from a Hilbert space $``$ determined up to the phase factor $`[|\psi >]`$ or, in other words, of one-dimensional projectors in $``$. Hence $`[|\psi >]`$ is the abstract class of unit vectors with respect to the following equivalence relation: $`|\psi >|\psi ^{}>`$ if $`|\psi >=e^{i\alpha }|\psi ^{}>`$ for some $`\alpha ๐`$. Let us notice that the scalar product of two distinct states is not well defined but its absolute value is. Also the one-dimensional projector $`|\psi ><\psi |`$ does not depend on the choice of a state vector $`|\psi >`$.
In order to define a subset of robust (completely stable) states let us notice that any environment-induced semigroup $`T_t`$ determines two linear closed and $`T_t`$-invariant subspaces $`\mathrm{Tr}()_{\mathrm{iso}}`$ and $`\mathrm{Tr}()_\mathrm{s}`$ in the Banach space of all trace class operators $`\mathrm{Tr}()`$. The subspace $`\mathrm{Tr}()_{\mathrm{iso}}`$ is called the isometric part and $`\mathrm{Tr}()_\mathrm{s}`$ the sweeping part. For the reader convenience we recall here some basic results of this isometric-sweeping decomposition. For proofs and a more detailed discussion see . The isometric and sweeping subspaces have the following properties:
a) $`\mathrm{Tr}()_{\mathrm{iso}}`$ and $`\mathrm{Tr}()_\mathrm{s}`$ are -invariant,
b) $`\mathrm{Tr}()_{\mathrm{iso}}\mathrm{Tr}()_\mathrm{s}`$ in the following sense: $`\varphi _1\mathrm{Tr}()_{\mathrm{iso}}\varphi _2\mathrm{Tr}()_\mathrm{s}`$ we have $`\mathrm{tr}\varphi _1\varphi _2=\mathrm{\hspace{0.17em}0}`$,
c) $`\mathrm{Tr}()=\mathrm{Tr}()_{\mathrm{iso}}\mathrm{Tr}()_\mathrm{s}`$, $`T_t=T_{1t}T_{2t}`$,
d) $`T_{1t}`$ is an invertible isometry given by a unitary group, i.e. $`T_{1t}\varphi =U_t\varphi U_t^{}`$ for any $`\varphi \mathrm{Tr}()_{\mathrm{iso}}`$,
e) $`T_{2t}`$ is sweeping, i.e. $`w^{}lim_t\mathrm{}T_{2t}\varphi =\mathrm{\hspace{0.17em}0}`$ for any $`\varphi \mathrm{Tr}()_\mathrm{s}`$, where $`w^{}`$ denotes the weak topology.
Hence, to any environment-induced semigroup corresponds a space of statistical states $`\mathrm{Tr}()_{\mathrm{iso}}`$ and associated with it an algebra of observables such that the evolution, when restricted to these spaces, is given by the Schrรถdinger unitary dynamics. In addition, $`\mathrm{Tr}()_{\mathrm{iso}}`$ has the following properties:
(i) if $`\varphi _1,\varphi _2\mathrm{Tr}()_{\mathrm{iso}}`$, then also $`\varphi _1\varphi _2\mathrm{Tr}()_{\mathrm{iso}}`$,
(ii) if projectors $`e,f\mathrm{Tr}()_{\mathrm{iso}}`$, then also $`ef\mathrm{Tr}()_{\mathrm{iso}}`$, where $`ef`$ denotes a projector onto the two-dimensional subspace spanned by the ranges of $`e`$ and $`f`$.
As a consequence, any one-dimensional projector $`e\mathrm{Tr}()_{\mathrm{iso}}`$ remains a projector during the evolution, and so $`S_{\mathrm{lin}}(T_te)=\mathrm{\hspace{0.17em}0}`$ for any $`t0`$. Therefore, we define a subset $`๐ฎ_0`$ of robust states by
$$๐ฎ_0=๐ฎ\mathrm{Tr}()_{\mathrm{iso}}$$
(1)
or, equivalently,
$$๐ฎ_0=\{e๐ฎ:S_{\mathrm{lin}}(T_te)=S_{\mathrm{lin}}(T_t^{}e)=\mathrm{\hspace{0.33em}0}\}$$
where $`T_t^{}`$ denotes the adjoint semigroup. If $`T_t^{}`$ commutes with $`T_t`$, then
$$\mathrm{tr}(T_t^{}e)^2=<T_t^{}e,T_t^{}e>_{HS}=<T_te,T_te>_{HS}=\mathrm{tr}(T_te)^2$$
where $`<,>_{HS}`$ is the scalar product in the Hilbert space of Hilbert-Schmidt operators, and so the condition $`S_{\mathrm{lin}}(T_te)=\mathrm{\hspace{0.17em}0}`$ implies that also $`S_{\mathrm{lin}}(T_t^{}e)=\mathrm{\hspace{0.17em}0}`$. Therefore, in such a case for a state $`e๐ฎ`$ to be robust it is enough that its linear entropy does not change in the course of evolution. It is worth noting that for quantum systems over finite dimensional Hilbert spaces it also turns out that the condition $`S_{\mathrm{lin}}(T_te)=\mathrm{\hspace{0.17em}0}`$ alone is sufficient for state $`e`$ to be in $`๐ฎ_0`$, see Appendix. If $`๐ฎ_0=๐ฎ`$, then the semigroup $`T_t`$ may be extended to a group of unitary automorphisms.
Obviously, any state from $`๐ฎ_0`$ will remain pure during the evolution and so remain in $`๐ฎ_0`$. Therefore, elements from $`๐ฎ_0`$ are the most probable candidates for โclassicalโ states. But the unitary evolution and thus perfect predictability alone does not suffice to accomplish our goal. Another feature distinguishing quantum from classical states, namely the validity of the superposition principle, has to be taken into account. In quantum mechanics it guarantees that any superposition of two distinct, and not necessarily orthogonal, states is again a legitimate quantum state. It means that for any pair of different one-dimensional projectors $`e_1`$ end $`e_2`$ we can associate a set of one-dimensional projectors $`e`$ given by $`e(e_1e_2)=e`$. Equivalently, we may write that
$$e=\frac{(z_1|\psi _1>+z_2|\psi _2>)(z_1^{}<\psi _1|+z_2^{}<\psi _2|)}{z_1|\psi _1>+z_2|\psi _2>^2}$$
where $`|\psi _1><\psi _1|=e_1`$, $`|\psi _2><\psi _2|=e_2`$ and $`z_1,z_2`$ are complex numbers. By contrast, classical states do not combine into another state. The only situation when their combination can be considered is inevitably tied to probability distributions on the phase space. Therefore, it is natural to assume that any non-trivial superposition of โclassicalโ states cannot be robust.
Definition 2.1. A state $`e๐ฎ`$ is called โclassicalโ if $`e๐ฎ_0`$ and for any $`f๐ฎ_0`$, $`fe`$, $`S(e,f)๐ฎ_0=\mathrm{}`$, where $`S(e,f)`$ denotes the collection of all states being non-trivial superpositions of $`e`$ and $`f`$. The collection of all โclassicalโ states we denote by $`๐ฎ_c`$.
Hence, although โclassicalโ states remain pure during the evolution, any of their superpositions deteriorates into a mixture. Under a mild, technical assumption namely that $`T_t`$ admits a holomorphic extension to a sector $`_ฯต=\{z:\mathrm{Re}z>0,|\mathrm{arg}z|<ฯต\}`$, for some $`ฯต>0`$, the loss of the purity of their superpositions happens instantaneously.
We are now in position to describe the structure of set $`๐ฎ_c`$.
Theorem 2.2. If $`๐ฎ_c\mathrm{}`$, then it consists of a family, possibly finite, of pairwise orthogonal states $`\{e_1,e_2,\mathrm{}\}`$ such that $`T_te_i=e_i`$ for all $`t0`$ and any index $`i`$.
Proof: Let $`e๐ฎ_c`$. We show that $`e`$ is orthogonal to any state $`f๐ฎ_0`$, $`fe`$. Suppose, on the contrary, that $`ef0`$. Then, because $`ef\mathrm{Tr}()_{\mathrm{iso}}`$, the state $`e^{}=efe`$ also belongs to $`\mathrm{Tr}()_{\mathrm{iso}}`$ and is orthogonal to $`e`$. All of these states can be considered as acting on a two-dimensional Hilbert space, the range of $`ef`$. Choosing an appropriate coordinate system we represent them by
$$e=\frac{1}{2}(I+\stackrel{}{n}_1\stackrel{}{\sigma }),f=\frac{1}{2}(I+\stackrel{}{n}_2\stackrel{}{\sigma })$$
with $`\stackrel{}{n}_1=(0,\mathrm{\hspace{0.17em}0},\mathrm{\hspace{0.17em}1})`$ and $`\stackrel{}{n}_2=(\mathrm{cos}\theta ,\mathrm{\hspace{0.17em}0},\mathrm{sin}\theta )`$, $`\theta [0,\pi /2)`$. Because the hermitian matrix $`i[e,f]\mathrm{Tr}()_{\mathrm{iso}}`$ is non-zero so its spectral projectors $`\frac{1}{2}(I\pm \stackrel{}{m}\stackrel{}{\sigma })`$, where $`\stackrel{}{m}=(0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}0})`$, also belong to $`\mathrm{Tr}()_{\mathrm{iso}}`$. On the other hand, they are superpositions of $`e`$ and $`e^{}`$. Therefore, $`S(e,e^{})๐ฎ_0\mathrm{}`$, what contradicts the assumption that $`e`$ is โclassicalโ. Hence $`ef`$ for any $`f๐ฎ_0`$, $`fe`$ and so $`๐ฎ_c=\{e_1,e_2,\mathrm{}\}`$ with $`e_ie_j=\delta _{ij}e_i`$. Finally, we show that $`T_te=e`$ for all $`t`$. If not so, then for any $`ฯต>0`$ we find an instant $`s`$ such that $`T_see`$ and $`T_see_1<ฯต`$. However, by the above argument, $`T_se`$ is orthogonal to $`e`$, so $`T_see_1=\mathrm{\hspace{0.17em}2}`$, the contradiction. $`\mathrm{}`$
Therefore, it turned out that โclassicalโ states, which are defined in a general way, form so-called pointer basis being introduced so far only on the operational level. Let us recall that pointer basis arises in a specific situation when before the measurement the quantum system was in an eigenstate of the measured observable. Such states are completely predictable since they do not evolve at all. By Theorem 2.2, they always correspond to points in a discrete classical phase space.
It is also clear that a unitary evolution $`T_t=e^{itH}e^{itH}`$ with $`H=H^{}`$, does not lead to the appearance of โclassicalโ states at all. Although $`๐ฎ_0=๐ฎ`$ in this case, $`๐ฎ_c=\mathrm{}`$ since any superposition of robust states is again robust. It is worth noting that, in general, even if โclassicalโ states exist, they may form an incomplete set of one-dimensional projectors.
3. โQuasi-classicalโ states
In this section we continue the investigation of states of an open quantum system which offer optimal predictability of their own future values. In the case when โclassicalโ states exist, they are the best candidates for states corresponding to classical points of phase space. If they are absent, it is natural to consider the states which are least affected by the interaction with the environment, that is, which are least prone to deteriorate into mixtures. Since linear entropy is a convenient measure of the loss of purity, we take its increase for initial states as a basic criterion. For a more complete analysis one should search for states which minimize the linear entropy over some finite period of time characteristic for the evolution of the system. To start with we define a quadratic form on the Hilbert space HS$`()`$ of Hilbert-Schmidt operators
$$B(\varphi )=<\varphi ,L(\varphi )>_{HS}$$
where $`\varphi D(L)\mathrm{Tr}()`$ and $`L`$ denotes the generator of semigroup $`T_t`$. Since $`\mathrm{Tr}()`$ is dense in HS$`()`$ so $`B`$ is densely defined. The closure of its symmetric part we denote by $`\lambda `$. By the Hille-Yosida theorem, the Lumer-Philips form, $`\lambda `$ is positive definite. It is clear that
$$\lambda (\varphi )=\frac{1}{2}\frac{d}{dt}S_{\mathrm{lin}}(T_t\varphi )|_{t=0}$$
(2)
whenever the corresponding derivative exists. Let
$$๐ฎ(a)=\{e๐ฎD(\lambda ):\lambda (e)=a\}$$
(3)
for $`a>0`$, and put $`a_0=inf\{a:๐ฎ(a)\mathrm{}\}`$. Guided by the previous considerations we define the set $`๐ฎ_s`$ of most stable states by $`๐ฎ_s=๐ฎ(a_0)`$, if it is non-empty. When $`๐ฎ(a_0)=\mathrm{}`$, then, in general, $`๐ฎ_s_{a<a_0+ฯต}๐ฎ(a)`$. The choice of $`ฯต`$ is somewhat arbitrary as it serves as the border between the preferred โquasi-classicalโ states and the โnon-classicalโ remainder. In this case, as was mentioned above, a further analysis examining the behavior of $`T_te`$ also for $`t>0`$ may be inevitable in order to select the set $`๐ฎ_s`$.
By combining predictability with the previously exploited principle expressing the fact that any superposition of two distinct preferred states cannot belong to the same class of stability we obtain the following.
Definition 3.1. A state $`e`$ is called โquasi-classicalโ if $`e๐ฎ_s`$ and for any $`f๐ฎ_s`$, $`fe`$, $`S(e,f)๐ฎ_s=\mathrm{}`$. The space of โquasi-classicalโ states will be denoted by $`๐ฎ_{qc}`$.
It should be pointed out that โquasi-classicalโ states can form an overcomplete set in contrast to the โclassicalโ states. On the other hand they may not exist at all. A simple example illustrating such a case is given by a dynamical semigroup on $`2\times 2`$ complex matrices with the following generator:
$$L(\rho )=i[H,\rho ]+(\mathrm{tr}\rho )I\mathrm{\hspace{0.33em}2}\rho $$
Then $`๐ฎ(a)=\mathrm{}`$ if $`a1`$ and $`๐ฎ(a)=๐ฎ`$ for $`a=\mathrm{\hspace{0.17em}1}`$. Hence $`๐ฎ_s`$ consists of all states and so any superposition of its two states again belongs to $`๐ฎ_s`$. Therefore $`๐ฎ_{qc}=\mathrm{}`$. In this case all states deteriorate into a completely mixed state in a uniform way.
4. Examples
Having discussed theoretical properties of โclassicalโ and โquasi-classicalโ states, let us now consider some physical examples.
4.1. Pointer states
Pointer states have been thoroughly discussed, see and references therein. They arise, for example, when the dynamical generator for the reduced density matrix is given by (see )
$$L(\rho )=i[H,\rho ]+\underset{i}{}P_i\rho P_i\frac{1}{2}\{P,\rho \}$$
where $`P_i`$ are one-dimensional orthogonal projectors, $`P=_iP_i`$, and Hamiltonian $`H`$ commutes with all $`P_i`$. Then $`๐ฎ_c=\{P_i\}`$.
4.2. Quantum Brownian motion
For quantum Brownian motion
$$\dot{\rho }=i[H,\rho ]D[x,[x,\rho ]]$$
(4)
which leads to an environment-induced semigroup, the rate of change of linear entropy for a state $`e=|\psi ><\psi |`$ is given by
$$\frac{d}{dt}S_{\mathrm{lin}}(T_te)|_{t=0}=\mathrm{\hspace{0.33em}4}D(<x^2><x>^2)$$
where $`<x>=<\psi |x|\psi >`$ and $`<x^2>=<\psi |x^2|\psi >`$. Hence $`\lambda (e)`$ is proportional to the dispersion in position of $`|\psi >`$ and so $`๐ฎ(a)\mathrm{}`$ for any $`a>0`$. However, $`๐ฎ(a=0)=\mathrm{}`$. A more detailed analysis shows that the space of the most stable states $`๐ฎ_s`$ consists of coherent states of the quantum harmonic oscillator . Because these states are represented by
$$|\alpha >=\mathrm{exp}(\frac{1}{2}|\alpha |^2)\underset{n=0}{\overset{\mathrm{}}{}}\frac{\alpha ^n}{(n!)^{1/2}}|n>$$
where $`\alpha `$ is a complex number and $`|n>`$ denotes the energy eigenstate, hence for any two distinct states $`|\alpha >`$ and $`|\beta >`$, $`\alpha \beta `$, none of their superpositions belongs to $`๐ฎ_s`$. Therefore, $`๐ฎ_sS(|\alpha ><\alpha |,|\beta ><\beta |)=\mathrm{}`$, and so the coherent states are โquasi-classicalโ. It is worth noting that coherent states of a harmonic oscillator coupled with up and down spins are also selected as the preferred states (โquasi-classicalโ in our terminology) in a model of the joint system of a spin-$`\frac{1}{2}`$ particle and a harmonic oscillator interacting with a zero-temperature bath of harmonic oscillators .
4.3. GRW spontaneous localization
Let us now examine the behavior of pure states $`e_\psi =|\psi ><\psi |`$ for a semigroup given by the master equation of the type discussed by Ghirardi, Rimini and Weber (see also )
$$\dot{\rho }=i[H,\rho ]+\underset{\mathrm{}}{\overset{\mathrm{}}{}}๐aG_a\rho G_a\kappa \rho $$
(5)
where $`G_a`$ is an operator of multiplication by a Gaussian function
$$g_a(x)=\kappa ^{1/2}(\frac{2\alpha }{\pi })^{1/4}e^{\alpha (xa)^2}$$
that is $`G_a\psi (x)=g_a(x)\psi (x)`$. Clearly, the above master equation leads to a dynamical semigroup which is also contractive in the operator norm. Hence $`\lambda `$ is well defined and
$$\lambda (e_\psi )=\kappa \underset{\mathrm{}}{\overset{\mathrm{}}{}}๐a(\mathrm{tr}G_ae_\psi )^2$$
$$=\kappa \underset{\mathrm{}}{\overset{\mathrm{}}{}}\underset{\mathrm{}}{\overset{\mathrm{}}{}}๐x๐y|\psi (x)|^2|\psi (y)|^2\underset{\mathrm{}}{\overset{\mathrm{}}{}}๐ag_a(x)g_a(y)$$
$$=\kappa [1\underset{\mathrm{}}{\overset{\mathrm{}}{}}\underset{\mathrm{}}{\overset{\mathrm{}}{}}๐x๐y|\psi (x)|^2|\psi (y)|^2e^{\alpha (xy)^2/2}]$$
Therefore, for any $`e_\psi `$, $`0<\lambda (e_\psi )<\kappa `$ and for any $`0<a<\kappa `$, $`๐ฎ(a)\mathrm{}`$. Clearly, the most stable states are those of Diracโs delta type, whereas states which are uniformly distributed over large intervals are strongly affected by the interaction.
4.4. Quantum stochastic process
This example shows that coherent states can be also selected as the โquasi-classicalโ states for quantum stochastic processes introduced by Davies . Quantum stochastic processes were introduced to describe rigorously certain continuous measurement processes. They can be constructed from two infinitesimal generators. The first is the generator $`Z`$ of a strongly continuous semigroup on a Hilbert space $``$, and the second is a stochastic kernel $`J`$, describing how the measuring apparatus interacts with the system. Let us recall that a stochastic kernel is a measure defined on the $`\sigma `$-algebra of Borel sets in some locally compact space and with values in the space of bounded positive linear operators on $`\mathrm{Tr}()`$. In this example we take Poincarรฉ disc $`D=\{\zeta ๐:|\zeta |<1\}`$ as the underlying topological space, and define
$$Z=iH\frac{\kappa }{2}\mathrm{๐}$$
where $`H`$ is the Hamiltonian of the system, $`\kappa >0`$ is the coupling constant. For $`ED`$ and $`\rho \mathrm{Tr}()`$ the stochastic kernel is defined by
$$\mathrm{tr}[J(E,\rho )A]=\kappa \underset{E}{}๐\mu (\zeta )\mathrm{tr}(e_\zeta \rho e_\zeta A)$$
where $`A`$ is a bounded linear operator on $``$, $`e_\zeta =|\zeta ><\zeta |`$ with $`|\zeta >`$ being a SU(1,1) coherent state, i.e. a holomorphic function on $`D`$
$$|\zeta >(z)=(1|\zeta |^2)(1z\zeta )^2$$
and
$$d\mu (\zeta )=\frac{1}{\pi }\frac{d\zeta d\overline{\zeta }}{(1|\zeta |^2)^2}$$
is a SU(1,1) invariant measure on $`D`$. In order to define a quantum stochastic process $`Z`$ and $`J`$ have to satisfy the following relation
$$\mathrm{tr}[J(D,e_\psi )]=2\mathrm{R}\mathrm{e}<\psi |Z|\psi >$$
$`e_\psi =|\psi ><\psi |`$, for all normalized vectors $`\psi D(Z)`$. It is straightforward to check that
$$\mathrm{tr}[J(D,e_\psi )]=\kappa \underset{Q}{}๐\mu (\zeta )\mathrm{tr}(e_\zeta e_\psi e_\zeta )=\kappa =2\mathrm{R}\mathrm{e}<\psi |Z|\psi >$$
The strongly continuous semigroup $`T_t`$ associated with the process is given by
$$T_t(\rho )=\mathrm{exp}(tZ^{})\rho \mathrm{exp}(tZ)+tJ(D,\rho )+o(t)$$
and so its generator reads
$$L(\rho )=i[H,\rho ]+\kappa \underset{Q}{}๐\mu (\zeta )e_\zeta \rho e_\zeta \kappa \rho $$
(6)
Obviously, it generates an environment-induced semigroup. It is worth noting that the integral formula above is a straightforward generalization of the von Neumann projection postulate to the case in which the family of states is overcomplete. Such a generator was thoroughly discussed in .
We now search for the most stable states with respect to semigroup $`T_t`$. Let us first note that no state is stable since, by Lemma 3.2 in , $`T_t`$ is strictly positive, that is $`T_te`$ is a faithful density matrix for every $`t>0`$ and any $`e๐ฎ`$. Hence $`๐ฎ_0=\mathrm{}`$ and so, in this case, there are no โclassicalโ states at all. However, the quadratic form $`\lambda `$ is bounded and allows to classify all states in the following way.
Proposition 4.1. $`\frac{2}{3}\kappa \lambda (e)<\kappa `$ for every $`e๐ฎ`$, and $`๐ฎ(a)\mathrm{}`$ if $`\frac{2}{3}\kappa a<\kappa `$. For $`a_0=\frac{2}{3}\kappa `$, $`๐ฎ(a_0)`$ consists exactly of the coherent states $`e_\zeta `$, $`|\zeta |<1`$.
Proof: Let $`e๐ฎ`$. Then, by definition,
$$\lambda (e)=\kappa [1\underset{Q}{}๐\mu (\zeta )(\mathrm{tr}ee_\zeta )^2]$$
Suppose $`e_n=|n><n|`$, where $`|n>(z)=\sqrt{n+\mathrm{\hspace{0.17em}1}}z^n`$, $`n๐\{0\}`$, is an orthonormal basis in $``$. It was shown in that
$$\underset{Q}{}๐\mu (\zeta )(\mathrm{tr}e_ne_\zeta )^2=\frac{n+\mathrm{\hspace{0.25em}1}}{(2n+\mathrm{\hspace{0.25em}1})(2n+\mathrm{\hspace{0.25em}3})}$$
Therefore,
$$0<\underset{Q}{}๐\mu (\zeta )(\mathrm{tr}ee_\zeta )^2\frac{1}{3}$$
for any $`e`$, and so $`\frac{2}{3}\kappa \lambda (e)<\kappa `$. Finally, notice that $`\lambda `$ is SU(1,1) invariant. Hence $`\lambda (e_0)=\lambda (\pi (g)e_0\pi (g)^{})=\frac{2}{3}\kappa `$ for any $`g`$SU(1,1). However, the set $`\{\pi (g)e_0\pi (g)^{}\}`$ coincides with the set of coherent states $`e_\zeta `$, $`|\zeta |<1`$. $`\mathrm{}`$
Hence, by definition, $`๐ฎ_s=๐ฎ(a_0)`$. Not surprisingly, the most stable states are the coherent ones. Next we show that they are โquasi-classicalโ. Because any $`|\zeta >`$ has the following representation
$$|\zeta >=(1|\zeta |^2)\underset{n=0}{\overset{\mathrm{}}{}}\sqrt{n+\mathrm{\hspace{0.25em}1}}\zeta ^n|n>$$
it follows that any superposition of two distinct coherent states $`|\zeta >`$ and $`|\zeta ^{}>`$ is not coherent, and thus $`๐ฎ_sS(e_\zeta ,e_\zeta ^{})=\mathrm{}`$. Hence the coherent states are โquasi-classicalโ and make up an analog of the pointer basis. They correspond to points in Poincarรฉ disc, which is the underlying space of the stochastic kernel representing the measuring apparatus.
In this section we have presented different types of examples which demonstrate how robust states can be selected and how non-stable states decohere to mixtures. Moreover, the rate of their deterioration to density matrices was examined. Such an analysis can be useful, for example, in quantum computing, where it is essential to control the process of decoherence in order to allow quantum bits to compute in parallel.
Acknowledgements
One of us (R.O.) would like to thank the A. von Humboldt Foundation for the financial support.
Appendix
Suppose dim$`<\mathrm{}`$ and let $`S_{\mathrm{lin}}(T_te)=\mathrm{\hspace{0.17em}0}`$ for some $`e๐ฎ`$. Then $`e๐ฎ_0`$.
Proof: Let $`P`$ be the projection onto $`\mathrm{Tr}()_{\mathrm{iso}}`$ along $`\mathrm{Tr}()_\mathrm{s}`$. Because $`P`$ extends to an orthogonal projection in the Hilbert space of Hilbert-Schmidt operators on $``$ so
$$T_te_2^2=P(T_te)_2^2+(idP)T_te_2^2$$
where $`_2`$ denotes the Hilbert-Schmidt norm. Since $`P`$ commutes with $`T_t`$ and $`T_te๐ฎ`$, we obtain that
$$1=T_t(Pe)_2^2+(idP)T_te_2^2$$
Because dim$`<\mathrm{}`$ so $`T_t`$ is relatively compact in the strong operator topology and so, by Theorem 24 in , $`lim_t\mathrm{}T_t\varphi P(T_t\varphi )_1=\mathrm{\hspace{0.17em}0}`$ for any trace class operator $`\varphi `$. Since $`_2_1`$ we obtain that $`lim_t\mathrm{}T_t(Pe)_2=\mathrm{\hspace{0.17em}1}`$. However, $`T_t`$ is also contractive in the norm $`_2`$, so $`Pe_2=\mathrm{\hspace{0.17em}1}`$ and hence $`Pe=e`$. It means that $`e\mathrm{Tr}()_{\mathrm{iso}}`$, and thus $`e๐ฎ_0`$. $`\mathrm{}`$
References
$`[1]`$ W.H. Zurek, Phys. Rev. D 24 (1981) 1516.
$`[2]`$ W.H. Zurek, Phys. Rev. D 26 (1982) 1862.
$`[3]`$ E. Joos and H.D. Zeh, Z. Phys. B 59 (1985) 223.
$`[4]`$ J.P. Paz and W.H. Zurek, Phys. Rev. D 48 (1993) 2728.
$`[5]`$ E. Joos, Decoherence through interaction with the environment, in: D. Giulini et al. (Eds.), Decoherence and the Appearance of a Classical World in Quantum Theory, Springer, Berlin, 1996.
$`[6]`$ D. Giulini et al. (Eds.), Decoherence and the Appearance of a Classical World in Quantum Theory, Springer, Berlin, 1996.
Ph. Blanchard et al. (Eds.), Decoherence: Theoretical, Experimental and Conceptual Problems, Lect. Notes Phys. 538, Springer, Berlin, 2000
$`[7]`$ W.H. Zurek, S. Habib and J.P. Paz, Phys. Rev. Lett. 70 (1993) 1187.
$`[8]`$ W.H. Zurek, Progr. Theor. Phys. 89 (1993) 281.
$`[9]`$ M.R. Gallis, Phys. Rev. A 53 (1996) 655.
$`[10]`$ R. Alicki and K. Lendi, Quantum Dynamical Semigroups and Applications, Lect. Notes Phys. 286, 1987.
$`[11]`$ B.L. Hu, J.P. Paz and Y. Zhang, Phys. Rev. D 45 (1992) 2843.
$`[12]`$ Ph. Blanchard and A. Jadczyk, Phys. Lett. A 175 (1993) 157.
$`[13]`$ Ph. Blanchard and A. Jadczyk, Phys. Lett. A 183 (1993) 272.
$`[14]`$ R. Olkiewicz, Rev. Math. Phys. 9 (1997) 719.
$`[15]`$ R. Olkiewicz, J. Math. Phys. 40 (1999) 1300.
$`[16]`$ W.G. Unruh and W.H. Zurek, Phys. Rev. D 40 (1989) 1071.
$`[17]`$ J. Twamley, Phys. Rev. D 48 (1993) 5730.
$`[18]`$ R. Olkiewicz, Commun. Math. Phys. 208 (1999) 245.
$`[19]`$ R. Olkiewicz, Structure of the algebra of effective observables in quantum mechanics, quanth-ph/0003032.
$`[20]`$ P. Carruthers and M.M. Nieto, Amer. J. Phys. 33 (1965) 537.
$`[21]`$ A. Venugopalan, Phys. Rev. A 61 (2000) 012102.
$`[22]`$ G.C. Ghirardi, A. Rimini and T. Weber, An attempt at a unified description of microscopic and macroscopic systems, in: V. Gorini and A. Frigerio (Eds.), Fundamental Aspects of Quantum Theory, NATO ASI Series B 144, Plenum Press, New York, 1986.
$`[23]`$ A. Jadczyk, Progr. Theor. Phys. 93 (1995) 631.
$`[24]`$ E.B. Davies, Commun. Math. Phys. 15 (1969) 277.
$`[25]`$ A.M. Perelomov, Commun. Math. Phys. 26 (1972) 222.
$`[26]`$ Ph. Blanchard and R. Olkiewicz, J. Stat. Phys. 94 (1999) 933. |
warning/0003/hep-ph0003314.html | ar5iv | text | # Bounds on extra quark-lepton generations from precision measurements
## 1 Introduction
The aim of this talk is to analyze to what extent precision measurements of IVB parameters allow to bound effectively the existence of extra chiral generations of heavy fermions, both quarks ($`q=U,D`$) and leptons ($`l=N,E`$). We will show that the case where *all* the extra particles are heavier than $`m_Z`$ is now excluded by more than $`2`$ standard deviations. However, if the masses of new neutrinos are assumed to be close to $`50\text{ GeV}`$, then additional generations become allowed and up to three extra families can exist within $`2.5\sigma `$. Finally, inclusion of new generations in SUSY extension of Standard Model is briefly discussed.
For simplicity, we will assume that the extra neutrinos are just ordinary massive Dirac particles, and that their mass is larger than $`45\text{ GeV}`$ so to avoid contradiction with the experimental data on the invisible $`Z`$ width. Also, we will neglect the possible mixing among new generations and the three existing ones, hence new fermions contribute to electroweak observables only through oblique corrections. On the other hand, we perform global fit of *all* precision data, studying both degenerate and non-degenerate extra generations on the equal footing. Taking the number of new generations $`N_g`$ as a continuous parameter, just as it was done with the determination of the number of neutrinos from invisible $`Z`$ width, we get a bound on it. The minimum $`\chi ^2`$ corresponds to $`N_g0.5`$, while the case $`N_g=1`$ is excluded by more than $`2`$ standard deviations.
## 2 Discussion
Heavy fermions. The comparison between theoretical predictions and experimental data is performed with the help of the computer code LEPTOP To simplify the analysis we start from the โhorizontally degenerateโ case $`(m_N=m_U)>(m_E=m_D=130\text{ GeV})`$ the lower bound on extra quark masses coming from Tevatron search, and in Fig. 2 we show the excluded domains in coordinates ($`N_g`$, $`\mathrm{\Delta }m(m_U^2m_D^2)^{1/2}`$). Any value of higgs mass above $`90\text{ GeV}`$ is allowed in our fits; minimum of $`\chi ^2`$ corresponds to $`N_g=0.5`$, and the case $`N_g=0`$ is within the $`1\sigma `$ domain. It is immediate to see that one extra generation is always excluded by more than $`2.5`$ standard deviations.
We checked that similar bounds are valid for the general choice of heavy masses of leptons and quarks. In particular we found that for the โcross degenerateโ case $`(m_E=m_U)>(m_N=m_D=130\text{ GeV}`$) one extra generation is excluded at $`2\sigma `$ level, while in the โanti-cross degenerateโ case $`(m_N=m_D)>(m_E=m_U=130\text{ GeV})`$ the limits are even stronger than in Fig. 2.
Light neutrinos. For particles with masses close to $`m_Z/2`$, oblique corrections drastically differ from what we have above $`m_Z`$. If we assume that the mixing angle between $`N`$ and the three known neutrinos is less than $`10^6`$, so to avoid the bound $`m_N>70รท80\text{ GeV}`$ from LEP II searches of the decays $`NlW^{}`$, we have that extra neutrinos as light as $`45\text{ GeV}`$ are still allowed by direct search experiments. In this case, effects of $`Z`$-boson wave function renormalization become relevant, and the quality of the fit can be even better than the SM. Analyzing all the electroweak observables, we conclude that presently a light neutral lepton $`N`$ cannot be excluded by precision measurements as well. As an example, in Fig. 2 we assume $`m_U=220\text{ GeV}`$, $`m_D=200\text{ GeV}`$, $`m_E=100\text{ GeV}`$ and draw the exclusion plot in coordinates ($`m_N,N_g`$): from this plot it is clear that for the case of extra generations with $`m_N50\text{ GeV}`$ even two new generations are allowed within $`1.5\sigma `$.
The case of SUSY. Concerning bounds on extra generations which occur in SUSY extensions, when superpartners are heavy their contributions to electroweak observables become power suppressed, and the same Standard Model exclusion plots shown in Fig. 2 and Fig. 2 are valid. The present lower bounds on the sparticle masses from direct searches leave mainly this decoupled domain. One possible exception is a contribution of the third generation squark doublet, enhanced by large stop-sbottom splitting. To see whether this contribution affects the bounds on extra generations found in the previous section, in Fig. 4 we analyze the simplest case of the absence of $`\stackrel{~}{t}_L\stackrel{~}{t}_R`$ mixing, setting the masses of all the extra fermions to the common value $`m_{N,E,U,D}=130\text{ GeV}`$ and showing the plot in coordinates $`(N_g,m_{\stackrel{~}{b}})`$. It is clear that even in the context of SUSY models new heavy generations are disfavored.
Situation qualitatively changes in the case of almost degenerate light chargino and neutralino. This possibility is yet not excluded โ dedicated search at LEP II by DELPHI still allows the existence of such particles with masses as low as $`45\text{ GeV}`$ if their mass difference is $`1\text{ GeV}`$ โ and even in this case radiative corrections are large. Fig. 4 demonstrates how the presence of a chargino-neutralino pair (dominated by higgsino) with mass $`57\text{ GeV}`$ relaxes the bounds shown on Fig. 2: we see that one extra generation of heavy fermions is now allowed within the $`1.5\sigma `$ domain.
## 3 Conclusions
Inclusion of new generations in Standard Model is not excluded by precision data if the new neutral leptons have masses close to $`m_Z/2`$ (see Fig. 2). In order to experimentally investigate this case a special search for the reaction $`e^+e^{}\gamma Z^{}\gamma N\overline{N}`$ with larger statistics and improved systematics is needed. Finally, further experimental search for light chargino and neutralino is of interest. These searches could close the existing windows of โlightโ extra particles, or open a door into a realm of New Physics.
## Acknowledgments
I wish to thank my collaborators V.A. Novikov, A.N. Rozanov, L.B. Okun and M.I. Vysotsky. This work was supported by DGICYT under grant PB98-0693 and by the TMR network grant ERBFMRX-CT96-0090 of the European Union. |
warning/0003/hep-ph0003305.html | ar5iv | text | # 1. Introduction
## 1. Introduction
For massless neutrinos, the leading contribution to low energy neutrino-photon scattering vanishes because the amplitude can be written in the form of a vector or axial-vector current coupling to the photons. The largest low energy interaction between photons and neutrinos is therefore $`\gamma \nu \gamma \gamma \nu `$ and the crossed channels. Rates for these processes have been calculated, for energies less than the electron mass, $`m_e`$, by using an effective action and, for larger energies, by a straight-forward calculation of box diagrams .
In the presence of magnetic fields neutrino photon elastic scattering does exist because the magnetic field gives an additional interaction โ it effectively replaces one of the photons in $`\gamma \nu \gamma \gamma \nu `$ with a zero momentum insertion. This too has been calculated; for low energies using the effective action for the $`23`$ processes and, at higher energies, by calculating triangle diagrams in the external field .
Our purpose here is to criticize mildly and extend slightly these external field results. In particular the effective action paper gave results for only one channel ($`\gamma \gamma \nu \overline{\nu }`$). We extend this to all channels with $`B`$ perpendicular and parallel for each channel. Results of the more general calculation of Ref. are given in three figures. We believe the curves in these figures are somewhat in error; for example, they sometimes disagree at low energies with our extended effective action results. Both references and express their results in terms of the critical magnetic field $`B_c=m_e^2/e=4.41\times 10^{13}`$ gauss. In each case the authors have taken $`e`$ to be in unrationalized units, $`e^2=1/137`$. whereas it should have been taken to be in natural units $`e^2=4\pi /137`$. Thus all their results are too large by a factor of $`4\pi `$.
## 2. Neutrino-photon cross sections in a uniform magnetic field
The calculation of Ref. uses the effective interaction given in Ref. . In this interaction, as in the Euler-Heisenberg Lagrangian upon which it is based, $`\alpha `$ equals $`e^2/4\pi `$ so the coupling of the magnetic field has been taken as this $`e`$. As discussed in Ref. the critical field, when written as $`m_e^2/e`$, uses $`e`$ in these same natural units. The value of the critical field, $`4.41\times 10^{13}`$ gauss, means that in natural (or rationalized) units 1 tesla $`=195\mathrm{eV}^2`$. The result of Ref. agrees in every respect with the result given below for $`\gamma \gamma \nu \overline{\nu }`$ with the external field perpendicular to the direction of the photons except in the replacement of $`B_c^2`$ by $`m_e^4/\alpha `$ where it should be replaced by $`m_e^4/4\pi \alpha `$. Since the curve $`\gamma \gamma \nu \overline{\nu }`$ in Fig. 2 of Ref. agrees, at low energy, with the result of Ref. these authors too must have made the wrong substitution for $`B_c`$.
### 2.1 Low energy interactions
The explicit calculations at low energy, using the effective action, are unremarkable and give the following results. For $`\gamma \gamma \nu \overline{\nu }`$ and $`\nu \overline{\nu }\gamma \gamma `$ the cross sections with center of mass energy $`\omega `$ and scattering angle $`z`$ are given by
$$\frac{d\sigma }{dz}=\sigma _0\frac{A}{4\pi }\left(\frac{\omega }{m_e}\right)^6\left(\frac{B}{B_c}\right)^2\left[2312\left(1N_1^2\right)+816N_1N_2zN_2^2(744+72z^2)\right],$$
(1)
where $`A=1`$ for $`\gamma \gamma \nu \overline{\nu }`$ and 1/2 for $`\nu \overline{\nu }\gamma \gamma `$. For electron neutrinos, including both the $`W`$ and $`Z`$-boson contributions, $`a=1/2+2\mathrm{sin}^2\theta _W`$ and
$$\sigma _0=\frac{G_F^2\alpha ^2a^2m_e^2}{(180)^2\pi ^2}=2.12\times 10^{54}\mathrm{cm}^2,$$
(2)
where $`G_F`$ is the Fermi constant and $`\theta _W`$ is the weak mixing angle. $`N_1`$ and $`N_2`$ depend on whether the $`B`$ field is parallel or perpendicular to the direction of the initial particles and are given in Table I.
The cross section for $`\gamma \nu \gamma \nu `$ is given by
$$\frac{d\sigma }{dz}=\frac{\sigma _0}{4\pi }\left(\frac{\omega }{m_e}\right)^6\left(\frac{B}{B_c}\right)^2\left[1211143z^2\left(N_1^2+N_2^2\right)(896+62z^2)+N_1N_2z(85911z^2)\right]$$
(3)
where terms odd in $`z`$ have been dropped. Using the $`N_1`$ and $`N_2`$ values and completing the integral over $`z`$ gives
$$\sigma =\frac{\sigma _0}{4\pi }\left(\frac{\omega }{m_e}\right)^6\left(\frac{B}{B_c}\right)^2๐ฟ,$$
(4)
where the $`๐ฟ`$ values for the six cases are given in Table I. Note that the result for $`\gamma \gamma \nu \overline{\nu }`$ with perpendicular $`B`$ agrees with Ref. if the $`4\pi `$ is ignored. Also note that the differential cross sections for parallel $`B`$ go as $`1z^2`$ as they must.
The $`๐ฟ`$ values show that the cross sections for $`\gamma \nu \gamma \nu `$ differ by a factor of about 4 depending on whether the magnetic field is parallel or perpendicular. The cross sections for $`\nu \overline{\nu }\gamma \gamma `$ differ by a factor of about 200. These differences are not present in the figures of Ref. .
### 2.2 Interactions at arbitrary center of mass energies
The authors of Refs. do a nice job of deriving the lowest order constant magnetic field contribution to photon-neutrino scattering, both from the correction to the electron propagator and from the phase factor. Their expression for the matrix element seems entirely correct with the exception of a relative minus sign in the tensor which multiplies the coefficient function $`C_{11}`$. The authors of Ref. inform us that this is a typographical error . We agree with all the coefficient functions $`C_1,\mathrm{},C_{11}`$.
The results of our calculation of neutrino-photon scattering in a constant magnetic field following the methods of Refs. and are shown in Figures 1, 2, and 3 for the three channels. All calculations are for $`B=0.1B_c`$ as is the case in Ref. . At low energies the cross sections agree almost exactly with Eq. (4). Note the relative factor of about 4 between the magnetic field directions for $`\gamma \nu \gamma \nu `$ and the similar factor of about 200 for $`\nu \overline{\nu }\gamma \gamma `$. Also note the infrared divergence at $`\omega =m_e`$ in the $`\gamma \gamma \nu \overline{\nu }`$ channels where the magnetic field interaction acts as a zero energy insertion on an external leg.
For large energy all of the cross sections grow as $`\omega ^2`$ except for $`\gamma \gamma \nu \overline{\nu }`$ with $`B`$ parallel to the photon direction, which decreases. After summation over photon polarizations, the cross section for this channel can be written as
$`\sigma `$ $`=`$ $`A_1F^{\mu \nu }F_{\mu \nu }+A_2(k_1^\mu F_{\mu \alpha }F^{\alpha \nu }k_{1\nu }+k_2^\mu F_{\mu \alpha }F^{\alpha \nu }k_{2\nu })`$ (5)
$`+A_3k_1^\mu F_{\mu \alpha }F^{\alpha \nu }k_{2\nu }+A_4(k_1^\mu F_{\mu \nu }k_2^\nu )^2,`$
where $`F^{\mu \nu }`$ is the field strength of the magnetic field and $`k_1`$ and $`k_2`$ are the momenta of the photons. The $`A_i`$ are combinations of the coefficient functions $`C_1,C_2,\mathrm{},C_{11}`$. For a magnetic field in a direction $`\stackrel{}{n}`$ this reduces to
$$\sigma 2A_1\omega ^2A_2[2(\widehat{k}_1\stackrel{}{n})^2(\widehat{k}_2\stackrel{}{n})^2]+\omega ^2A_3[1+\widehat{k}_1\stackrel{}{n}\widehat{k}_2\stackrel{}{n}].$$
(6)
Thus for $`\gamma \gamma \nu \overline{\nu }`$ with $`\stackrel{}{n}`$ parallel to $`\widehat{k}_1`$ and $`\widehat{k}_2`$ the coefficients of $`A_2`$ and $`A_3`$ vanish. $`A_1`$, which in this channel depends only on $`C_1`$, falls as $`\mathrm{ln}^2(\omega ^2/m_e^2)/\omega ^2`$ for $`\omega >>m_e`$, giving the decrease at large $`\omega `$ seen in Fig. (1). $`A_2`$ and $`A_3`$ do not vanish for $`\stackrel{}{n}`$ perpendicular to $`\widehat{k}_1`$ and $`\widehat{k}_2`$ nor do they vanish for either $`B`$ direction in the other channels.
## 3. Conclusions
We have extended the low energy effective Lagrangian calculation of Ref. to include all channels and directions of the magnetic field B. In Figs. (1-3), we have given correct values for the cross sections at all energies, which, while shown for $`B=0.1B_c`$, can be scaled by $`B^2`$. (Of course, following Ref. and , we have essentially done perturbation theory in $`B/B_c`$ so our results are not valid for much larger $`B`$.) All calculations were performed for electron neutrinos but this too can be easily changed by using $`a=1/22\mathrm{sin}^2\theta _W`$ for the other neutrino types.
It turns out that neutrino photon scattering in an external magnetic field is less important, relative to the $`23`$ processes, than previously thought . Since the effects of these processes on stars was already calculated to be small we have not repeated the determination of the stellar energy loss rates or mean free paths.
## Acknowledgement
We would like thank Chung Kao, Vic Teplitz and Roberto Vega for helpful conversations and R. Shaisultanov and G.-L. Lin for comments. This work was supported in part by the National Science Foundation under grant PHY-9802439 and by the Department of Energy under Grant No. DE-FG13-93ER40757.
## References
## Tables
## Figures |
warning/0003/math0003110.html | ar5iv | text | # I. Introduction
## I. Introduction
It is known that quantum deformation of Lie group $`GL(2)`$ with central quantum determinant is classified into two types : the standard deformation $`GL_q(2)`$ and the Jordanian deformation $`GL_h(2)`$ . The representation theory of $`GL_q(2)`$ has been studied extensively and we know that its contents are quite rich (See, for instance, Refs. ). On the other hand, the representation theory of $`GL_h(2)`$ has not been developed yet. There are some works studying differential geometry on quantum $`h`$-plane and on $`SL_h(2)`$ itself . However, the representation functions for $`GL_h(2)`$, the most basic ingredient of representation theories, has not been known. Recently Chakrabarti and Quesne showed that the representation functions for two-parametric extension of $`GL_h(2)`$ can be obtained from the standard deformed ones via a contraction method and gave explicit form of representation functions for some low dimensional cases. In Ref., the present author shows that the Jordanian deformation of symplecton for $`sl(2)`$ gives a natural basis for a representation of $`SL_h(2)`$ and he also gives another basis in terms of quantum $`h`$-plane.
The purpose of the present paper is to obtain explicit formulae for $`SL_h(2)`$ representation functions using the tensor operator technique and to investigate their properties. Representation functions are also called Wignerโs $`D`$-functions in physicistโs terminology. We use both terms and concentrate ourselves to the finite dimensional highest weight irreducible representations of $`SL_h(2)`$ throughout the present paper. In order to make a comparison between $`D`$-functions for $`SL_q(2)`$ and $`SL_h(2)`$, let us recall some known properties of $`D`$-functions for $`SL_q(2)`$ : (a) Wignerโs product law , (b) recurrence relations , (c) orthogonality (d) RTT type relations , (e) $`D`$-functions can be written in terms of the little $`q`$-Jacobi polynomials , (f) generating function . We will show, in this paper, that many of them have counterparts in the representation theory of $`SL_h(2)`$. Only exception is the generating function, it is not presented in this paper. Of course it does not mean that the generating function for the $`D`$-functions of $`SL_h(2)`$ dose not exist.
The plan of this paper is as follows: we present the definitions of $`SL_h(2)`$ and its dual quantum algebra $`๐ฐ_h(sl(2))`$ in the next section. In ยงIII, before deriving the explicit formulae for the representation functions, we discuss general features of them which are valid for any kind of deformation of $`SL(2)`$ under the assumption that the representation theory of the dual quantum algebra has a one-to-one correspondence with the undeformed $`sl(2).`$ Then we shall write down the recurrence relations for $`SL_h(2)D`$-functions. ยงIV is a brief review of the $`D`$-functions for Lie group $`SL(2)`$ (and $`GL(2)`$). We emphasize that the $`D`$-functions for $`GL(2)`$ form, in a certain boson realization, irreducible tensor operators of the Lie algebra $`gl(2)gl(2).`$ In ยงV, a tensor operator technique is used to obtain the boson realization of the generators of the Jordanian quantum group $`GL_h(2)`$, then it is generalized to obtain the $`D`$-functions for $`GL_h(2)`$. We shall apply the same technique to show that the $`D`$-functions for $`SL_h(2)`$ can be expressed in terms of Jacobi polynomials. This method will be applied to obtain a boson realization for two-parametric extension of the Jordanian deformation of $`GL(2)`$ in ยงVI. ยงVII is concluding remarks.
## II. $`๐บ๐ณ_๐\mathbf{(}\mathrm{๐}\mathbf{)}`$ and its Dual
The Jordanian quantum group $`GL_h(2)`$ is generated by four elements $`x,y,u`$ and $`v`$ subject to the relations
$`[v,x]=hv^2,[u,x]=h(Dx^2),`$
$`[v,y]=hv^2,[u,y]=h(Dy^2),`$ (2.1)
$`[x,y]=h(xvyv),[v,u]=h(xv+vy),`$
where $`D=xyuvhxv`$ is the quantum determinant generating the center of $`GL_h(2)`$. This is a Hopf algebra and Hopf algebra mappings have a similar form as $`GL_q(2)`$. However, explicit form of the mappings is not necessary in the following discussion. By setting $`D=1`$, we obtain $`SL_h(2)`$ from $`GL_h(2).`$
The quantum algebra dual to $`GL_h(2)`$ is denoted by $`๐ฐ_h(gl(2))`$, and defined by the same commutation relations as the Lie algebra $`gl(2)`$
$$[J_0,J_\pm ]=\pm 2J_\pm ,[J_+,J_{}]=J_0,[Z,]=0.$$
(2.2)
However, their Hopf algebra mappings are modified via twisting by the invertible element $`๐ฐ_h(gl(2))^2`$
$$=\mathrm{exp}\left(\frac{1}{2}J_0\sigma \right),\sigma =\mathrm{ln}(12hJ_+).$$
(2.3)
The coproduct $`\mathrm{\Delta }`$, counit $`ฯต`$ and antipode $`S`$ for $`๐ฐ_h(gl(2))`$ are obtained from those for $`gl(2)`$ by
$$\mathrm{\Delta }=\mathrm{\Delta }_0^1,ฯต=ฯต_0,S=\mu S_0\mu ^1,$$
(2.4)
where the mappings with suffix $`0`$ stand for the Hopf algebra mappings for $`gl(2)`$. The elements $`\mu `$ and $`\mu ^1`$ are defined, using the product $`m`$ for $`gl(2)`$, by
$$\mu =m(idS_0)(),\mu ^1=m(S_0id)(^1).$$
(2.5)
The twist element $``$ is not depend on the central element $`Z`$ so that the Hopf algebra mappings for $`Z`$ remain undeformed. Therefore the Jordanian quantum algebra obtained by the twist element (2.3) has the decomposition $`๐ฐ_h(gl(2))=๐ฐ_h(sl(2))u(1)`$. The Jordanian quantum algebra $`๐ฐ_h(gl(2))`$ is a triangular Hopf algebra whose universal $`R`$-matrix is given by $`=_{12}^1`$.
It is obvious, from the commutation relation (2.2), that $`๐ฐ_h(gl(2))`$ and $`gl(2)`$ have the same finite dimensional highest weight irreducible representations. Furthermore we can easily see that tensor product of two irreducible representations (irreps) is completely reducible and decomposed into irreps in the same way as $`gl(2),`$ since the Clebsch-Gordan coefficients (CGC) for $`๐ฐ_h(gl(2))`$ are product of the ones for $`gl(2)`$ and matrix elements of the twist element $``$. For the $`๐ฐ_h(sl(2))`$ sector, this is carried out in Ref.. The CGC for $`๐ฐ_h(sl(2))`$ in another basis are discussed in Ref.
Let $`\mathrm{\Delta },ฯต`$ be the coproduct and counit for $`GL_h(2)`$, respectively. We use the same notations for the Hopf algebra mappings of both $`GL_h(2)`$ and $`๐ฐ_h(gl(2))`$, however, this may not cause serious confusion. A vector space (representation space) $`V`$ is called right $`GL_h(2)`$ comodule, if there exist a map $`\rho :VVGL_h(2)`$ such that the following relations are satisfied
$$(\rho id)\rho =(id_V\mathrm{\Delta })\rho ,(id_Vฯต)\rho =id_V,$$
(2.6)
where $`id_V`$ stands for the identity map in $`V`$. The left comodule is defined in the similar manner. Using the bases $`\{e_i|i=1,2,\mathrm{},n\}`$ of $`V`$, the map $`\rho `$ is written as
$$\rho (e_i)=\underset{j}{}e_j๐_{ji},$$
(2.7)
it follows that the relation (2.6) are rewritten as
$$\mathrm{\Delta }(๐_{ij})=\underset{k}{}๐_{ik}๐_{kj},ฯต(๐_{ij})=\delta _{ij}.$$
(2.8)
We call $`๐_{ij}GL_h(2)`$ satisfying (2.7) and (2.8) the $`D`$-function for $`GL_h(2)`$.
## III. Properties of $`D`$-functions
### A. Wignerโs Product Law and RTT Type Relations
Before deriving the explicit formulae for $`SL_h(2)`$ $`D`$-functions , one can discuss some important properties of $`D`$-functions such as Wignerโs product law, recurrence relations, RTT type relations and so on, using the definition of universal $`T`$-matrix . The explicit expression of the universal $`T`$-matrix is not necessary. The universal $`T`$-matrix for the standard deformation of $`GL(2)`$ is given in Ref., while it is not known for the Jordanian deformation of $`GL(2)`$.
The discussion in this subsection is quite general. We shall present it so as to be applicable to any kind of deformation of $`SL(2)`$ (standard, Jordanian, two-parametric extension, anything else if any). Then we will write down the results explicitly for the Jordanian deformation of $`SL(2)`$ in the next subsection. It will also be seen that the discussion is easily extended to other groups.
Let $`๐ข`$ and $`๐ `$ be deformation of Lie group $`SL(2)`$ and Lie algebra $`sl(2)`$, respectively. The duality between $`๐ข`$ and $`๐ `$ are expressed, by choosing suitable bases, in terms of the universal $`T`$-matrix . Let $`x^\alpha `$ and $`X_\alpha `$ be elements of a basis of $`๐ข`$ and $`๐ `$, respectively. They are chosen as follows: the product is given by
$$x^\alpha x^\beta =\underset{\gamma }{}h_\gamma ^{\alpha ,\beta }x^\gamma ,X_\alpha X_\beta =\underset{\gamma }{}f_{\alpha ,\beta }^\gamma X_\gamma ,$$
(3.1)
the coproduct is given by
$$\mathrm{\Delta }(x^\alpha )=\underset{\beta ,\gamma }{}f_{\beta ,\gamma }^\alpha x^\beta x^\gamma ,\mathrm{\Delta }(X_\alpha )=\underset{\beta ,\gamma }{}h_\alpha ^{\beta ,\gamma }X_\beta X_\gamma .$$
(3.2)
Then the universal $`T`$-matrix $`๐ฏ`$ is defined by
$$๐ฏ=\underset{\alpha }{}x^\alpha X_\alpha .$$
(3.3)
We assume that the deformed algebra $`๐ `$ has the same finite dimensional highest weight irreps as $`sl(2)`$, that is, (1) each irrep is classified by the spin $`j`$ and a irrep basis $`|jm`$ is specified by $`j`$ and the magnetic quantum number $`m`$, (2) tensor product of irreps $`j_1`$ and $`j_2`$ is completely reducible
$$j_1j_2=j_1+j_2j_1+j_21\mathrm{}|j_1j_2|.$$
We further assume that vectors $`|jm`$ are complete and orthonormal. Then the $`D`$-functions for $`๐ข`$ is obtained by
$$๐_{m^{},m}^j=jm^{}\left|๐ฏ\right|jm=\underset{\alpha }{}x^\alpha jm^{}\left|X_\alpha \right|jm.$$
(3.4)
For the standard two-parametric deformation of $`GL(2)`$, the RHS of (3.4) was computed and it was shown that (3.4) coinceided with the $`D`$-functions obtained by another method . In our case, we show that the $`D`$-functions (3.4) satisfy (2.8) by making use of the relations (3.1) and (3.2). The coproduct of $`๐_{m^{},m}^j`$ is computed as
$`\mathrm{\Delta }(๐_{m^{},m}^j)`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}\mathrm{\Delta }(x^\alpha )jm^{}\left|X_\alpha \right|jm={\displaystyle \underset{\beta ,\gamma }{}}x^\beta x^\gamma jm^{}\left|X_\beta X_\gamma \right|jm`$
$`=`$ $`{\displaystyle \underset{\beta ,\gamma ,k}{}}x^\beta x^\gamma jm^{}\left|X_\beta \right|jkjk\left|X_\gamma \right|jm={\displaystyle \underset{k}{}}๐_{m^{},k}^j๐_{k,m}^j.`$
To compute the counit for $`๐_{m^{},m}^j`$, we use the identiy obtained from the definition of counit
$$\underset{\beta ,\gamma }{}f_{\beta ,\gamma }^\alpha ฯต(x^\beta )x^\gamma =x^\alpha .$$
(3.5)
Using this relation, the universal $`T`$-matrix is rewritten as
$`๐ฏ`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}x^\alpha X_\alpha ={\displaystyle f_{\beta ,\gamma }^\alpha ฯต(x^\beta )x^\gamma X_\alpha }={\displaystyle ฯต(x^\beta )x^\gamma X_\beta X_\gamma }`$
$`=`$ $`({\displaystyle \underset{\beta }{}}ฯต(x^\beta )X_\beta )๐ฏ.`$
It follows that
$$(\underset{\beta }{}ฯต(x^\beta )X_\beta )=(ฯตid)(๐ฏ)=1.$$
(3.6)
Therefore the counit for $`D`$-functions is
$$ฯต(๐_{m^{},m}^j)=jm^{}\left|(ฯตid)(๐ฏ)\right|jm=jm^{}|jm=\delta _{m^{},m}.$$
We first show that the $`D`$-functions (3.4) satisfy the analogous relations to Wignerโs product law. Let us denote the CGC for $`๐ `$ by $`\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}`$, $`i.e.`$,
$$|(j_1j_2)jm=\underset{m_1,m_2}{}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}|j_1m_1|j_2m_2.$$
(3.7)
We write the inverse of the above relation as follows:
$$|j_1m_1|j_2m_2=\underset{j,m}{}\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}|(j_1j_2)jm.$$
(3.8)
Then an analogue of Wignerโs product law reads
###### Theorem 3.1
The $`D`$-functions for $`๐ข`$ satisfy the relation
$$\delta _{j,j^{}}๐_{m^{},m}^j=\underset{k_1,k_2,m_1,m_2}{}\mathrm{}_{k_1,k_2,m^{}}^{j_1,j_2,j^{}}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2}.$$
(3.9)
$`Proof`$ : Because of the relations (3.1) and (3.2), one can show that
$$(id\mathrm{\Delta })(๐ฏ)=\underset{\alpha ,\beta }{}x^\alpha x^\beta X_\alpha X_\beta ,(\mathrm{\Delta }id)(๐ฏ)=\underset{\alpha ,\beta }{}x^\alpha x^\beta X_\alpha X_\beta .$$
(3.10)
It follows that
$$(id\mathrm{\Delta })(๐ฏ)|(j_1j_2)jm=\underset{\alpha ,\beta ,m_1,m_2}{}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}x^\alpha x^\beta X_\alpha |j_1m_1X_\beta |j_2m_2.$$
(3.11)
The LHS of (3.11) is rewritten as
$$\underset{m^{}}{}๐_{m^{},m}^j|(j_1j_2)jm^{}=\underset{m^{},k_1,k_2}{}\mathrm{\Omega }_{k_1,k_2,m^{}}^{j_1,j_2,j}๐_{m^{},m}^j|j_1k_1|j_2k_2.$$
The RHS of (3.11) is rewritten as
$`{\displaystyle \mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}j_1k_1\left|X_\alpha \right|j_1m_1j_2k_2\left|X_\beta \right|j_2m_2x^\alpha x^\beta |j_1k_1|j_2k_2}`$
$`=`$ $`{\displaystyle \mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2}|j_1k_1|j_2k_2}.`$
Thus we obtain
$$\underset{m^{}}{}\mathrm{\Omega }_{k_1,k_2,m^{}}^{j_1,j_2,j}๐_{m^{},m}^j=\underset{m_1,m_2}{}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2}.$$
(3.12)
Using the orthogonality of $`\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}`$ and $`\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}`$, the theorem is proved. $`\mathrm{}`$
###### Corollary 3.2
The $`D`$-functions also satisfy the following relations
$`{\displaystyle \underset{m^{}}{}}\mathrm{\Omega }_{k_1,k_2,m^{}}^{j_1,j_2,j}๐_{m^{},m}^j={\displaystyle \underset{m_1,m_2}{}}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2},`$ (3.13)
$`{\displaystyle \underset{m}{}}\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}๐_{m^{},m}^j={\displaystyle \underset{k_1,k_2}{}}\mathrm{}_{k_1,k_2,m^{}}^{j_1,j_2,j}๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2},`$ (3.14)
$`๐_{k_1,m_1}^{j_1}๐_{k_2,m_2}^{j_2}={\displaystyle \underset{j,m,m^{}}{}}\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}\mathrm{\Omega }_{k_1,k_2,m^{}}^{j_1,j_2,j}๐_{m^{},m}^j.`$ (3.15)
$`Proof`$ : (3.13) has already been obtained in the proof of Theorem 3.9, see (3.12). Others can be obtained from (3.13) by the orthogonality of $`\mathrm{\Omega }`$ and $`\mathrm{}`$. $`\mathrm{}`$
For $`๐ข=SL_q(2)`$ and $`๐ =๐ฐ_q(sl(2))`$, the CGC $`\mathrm{\Omega },\mathrm{}`$ are given by the $`q`$-analogue of the CGC of $`sl(2)`$ : $`\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}=\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}={}_{q}{}^{}C_{m_1,m_2,m}^{j_1,j_2,j}`$ . From the relations (3.9) and (3.13) - (3.15), the recurrence relations and the orthogonality of $`SL_q(2)`$ $`D`$-functions are obtained .
Next we show that the $`D`$-functions (3.4) satisfy the RTT type relation.
###### Theorem 3.3
The $`D`$-functions for $`๐ข`$ satisfy
$$\underset{s_1,s_2}{}(R^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}๐_{s_1,k_1}^{j_1}๐_{s_2,k_2}^{j_2}=\underset{s_1,s_2}{}๐_{m_2,s_2}^{j_2}๐_{m_1,s_1}^{j_1}(R^{j_1,j_2})_{s_1,s_2}^{k_1,k_2},$$
(3.16)
where $`(R^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}`$ are the matrix elements of the universal $`R`$-matrix for $`๐ `$
$$(R^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}=j_1m_1|j_2m_2\left|\right|j_1s_1|j_2s_2.$$
$`Remark`$ : For $`j_1=j_2=1/2`$, the matrix elements for $``$ are evaluated in the fundamental representation of $`๐ `$. Therefore, the relation (3.16) is reduced to the defining relation of $`๐ข`$ in FRT-formalism . This implies that $`๐_{m^{},m}^{\frac{1}{2}}`$ are generators of $`๐ข.`$
$`Proof`$ : The relation (3.16) can be proved by evaluating matrix elements of the RTT type relation for the universal $`T`$-matrix . We define
$$๐ฏ_1=x^\alpha X_\alpha 1,๐ฏ_2=x^\alpha 1X_\alpha ,$$
then
$`๐ฏ_1๐ฏ_2={\displaystyle \underset{\alpha ,\beta }{}}x^\alpha x^\beta X_\alpha X_\beta ={\displaystyle \underset{\alpha }{}}x^\alpha \mathrm{\Delta }(X_\alpha ),`$
$`๐ฏ_2๐ฏ_1={\displaystyle \underset{\alpha ,\beta }{}}x^\beta x^\alpha X_\alpha X_\beta ={\displaystyle \underset{\alpha }{}}x^\alpha \mathrm{\Delta }^{}(X_\alpha ),`$
where $`\mathrm{\Delta }^{}`$ stands for the opposite coproduct. It follows that
$$๐ฏ_2๐ฏ_1=\underset{\alpha }{}x^\alpha \mathrm{\Delta }(X_\alpha )^1,$$
thus we obtain
$$(1)๐ฏ_1๐ฏ_2=๐ฏ_2๐ฏ_1(1).$$
Evaluating the matrix elements on $`1|j_1k_1|j_2k_2`$, the theorem is proved. $`\mathrm{}`$
For $`๐ข=SL_q(2)`$, the relaiton (3.16) was proved by Nomura . However, Theorem 3.3 shows that the relation (3.16) holds for any kind of deformation of $`SL(2)`$. In Ref., the $`D`$-functions for $`SL_q(2)`$ are interpreted as the wave functions of quantum symmetric tops in noncommutative space.
### B. Recurrence Relations and Orthogonality-like Relations
In this subsection, the recurrence relations and the orthogonality-like relations of the $`SL_h(2)`$ $`D`$-functions are derived as a consequence of the theorems in the previous subsection. It is known that the CGC for $`๐ฐ_h(sl(2))`$ are given in terms of the CGC for $`sl(2)`$ and the matrix elements of the twist element $``$
$$\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}=\underset{s_1,s_2}{}C_{s_1,s_2,m}^{j_1,j_2,j}(F^{j_1,j_2})_{m_1,m_2}^{s_1,s_2},$$
(3.17)
where $`C_{s_1,s_2,m}^{j_1,j_2,j}`$ is the CGC for $`sl(2)`$ and $`(F^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}`$ is given by
$$(F^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}=j_1,m_1|j_2,m_2\left|\right|j_1,s_1|j_2,s_2.$$
The explicit formula for $`(F^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}`$ and the next relation are found in Ref..
$$(F^{j_1,j_2})_{m_1,m_2}^{s_1,s_2}=((F^1)^{j_1,j_2})_{s_1,s_2}^{m_1,m_2}.$$
(3.18)
The CGC for $`๐ฐ_h(sl(2))`$ satisfy the orthogonality relations because of
$$\mathrm{}_{m_1,m_2,m}^{j_1,j_2,j}=(1)^{j_1+j_2j}\mathrm{\Omega }_{m_1,m_2,m}^{j_1,j_2,j}=\underset{s_1,s_2}{}C_{s_1,s_2,m}^{j_1,j_2,j}((F^1)^{j_1,j_2})_{s_1,s_2}^{m_1,m_2}.$$
(3.19)
The relation(3.18) and the well known property of the $`sl(2)`$ CGC are used in the last equality.
Note that we have known the following fact because of the remark to Theorem 3.3.
###### Proposition 3.4
$`๐_{m^{},m}^{\frac{1}{2}}`$ are the generators of $`SL_h(2)`$
$$\left(\begin{array}{cc}๐_{\frac{1}{2},\frac{1}{2}}^{\frac{1}{2}}& ๐_{\frac{1}{2},\frac{1}{2}}^{\frac{1}{2}}\\ ๐_{\frac{1}{2},\frac{1}{2}}^{\frac{1}{2}}& ๐_{\frac{1}{2},\frac{1}{2}}^{\frac{1}{2}}\end{array}\right)=\left(\begin{array}{cc}x& u\\ v& y\end{array}\right)_.$$
(3.20)
Let us consider the case that $`j_1`$ is arbitrary and $`j_2=1/2`$ in order to derive the recurrence relations for $`SL_h(2)`$ $`D`$-functions. In this case, the $`F`$-coefficients have a simple form
$`(F^{j_1,\frac{1}{2}})_{k_1,k_2}^{m_1,\frac{1}{2}}=\delta _{k_1,m_1}\delta _{k_2,\frac{1}{2}},`$
$`(F^{j_1,\frac{1}{2}})_{k_1,k_2}^{m_1,\frac{1}{2}}=\delta _{k_1,m_1}(\delta _{k_2,\frac{1}{2}}2m_1h\delta _{k_2,\frac{1}{2}}),`$
$`((F^1)^{j_1,\frac{1}{2}})_{m_1,\frac{1}{2}}^{n_1,n_2}=\delta _{m_1,n_1}(\delta _{n_2,\frac{1}{2}}+2m_1h\delta _{n_2,\frac{1}{2}}),`$
$`((F^1)^{j_1,\frac{1}{2}})_{m_1,\frac{1}{2}}^{n_1,n_2}=\delta _{m_1,n_1}\delta _{n_2,\frac{1}{2}}.`$
One can use Wingerโs product law, expressed in the form of (3.13) and (3.14), to derive the recurrence relations for $`๐_{m^{},m}^j`$ which are reduced to the known recurrence relations of the $`SL(2)`$ $`D`$-functions in the limit of $`h=0.`$
###### Proposition 3.5
The $`SL_h(2)`$ $`D`$-functions satisfy the following recurrence relations
| (i) | $`\sqrt{j+k}๐_{k,m}^j(2k1)h\sqrt{jk+1}๐_{k1,m}^j`$ |
| --- | --- |
| | $`=\sqrt{j+m}๐_{k\frac{1}{2},m\frac{1}{2}}^{j\frac{1}{2}}x+\sqrt{jm}๐_{k\frac{1}{2},m+\frac{1}{2}}^{j\frac{1}{2}}(u(2m+1)hx),`$ |
| (ii) | $`\sqrt{jk}๐_{k,m}^j=\sqrt{j+m}๐_{k+\frac{1}{2},m\frac{1}{2}}^{j\frac{1}{2}}v+\sqrt{jm}๐_{k+\frac{1}{2},m+\frac{1}{2}}^{j\frac{1}{2}}(y(2m+1)hv),`$ |
| (iii) | $`\sqrt{j+n}๐_{m,n}^j=\sqrt{j+m}๐_{m\frac{1}{2},n\frac{1}{2}}^{j\frac{1}{2}}(x+(2m1)hv)+\sqrt{jm}๐_{m+\frac{1}{2},n\frac{1}{2}}^{j\frac{1}{2}}v,`$ |
| (iv) | $`\sqrt{jn}๐_{m,n}^j+\sqrt{j+n}(2n+1)h๐_{m,n+1}^j`$ |
| | $`=\sqrt{j+m}๐_{m\frac{1}{2},n+\frac{1}{2}}^{j\frac{1}{2}}(u+(2m1)hy)+\sqrt{jm}๐_{m+\frac{1}{2},n+\frac{1}{2}}^{j\frac{1}{2}}y,`$ |
| (v) | $`\sqrt{jk+1}๐_{k,m}^j+(2k1)h\sqrt{j+k}๐_{k1,m}^j`$ |
| | $`=\sqrt{jm+1}๐_{k\frac{1}{2},m\frac{1}{2}}^{j+\frac{1}{2}}x\sqrt{j+m+1}๐_{k\frac{1}{2},m+\frac{1}{2}}^{j+\frac{1}{2}}(u(2m+1)hx),`$ |
| (vi) | $`\sqrt{j+k+1}๐_{k,m}^j=\sqrt{jm+1}๐_{k+\frac{1}{2},m\frac{1}{2}}^{j+\frac{1}{2}}v`$ |
| | $`+\sqrt{j+m+1}๐_{k+\frac{1}{2},m+\frac{1}{2}}^{j+\frac{1}{2}}(y(2m+1)hv),`$ |
| (vii) | $`\sqrt{jn+1}๐_{m,n}^j=\sqrt{jm+1}๐_{m\frac{1}{2},n\frac{1}{2}}^{j+\frac{1}{2}}(x+(2m1)hv)`$ |
| | $`\sqrt{j+m+1}๐_{m+\frac{1}{2},n\frac{1}{2}}^{j+\frac{1}{2}}v,`$ |
| (viii) | $`\sqrt{j+n+1}๐_{m,n}^j\sqrt{jn}(2n+1)h๐_{m,n+1}^j`$ |
| | $`=\sqrt{jm+1}๐_{m\frac{1}{2},n+\frac{1}{2}}^{j+\frac{1}{2}}(u+(2m1)hy)+\sqrt{j+m+1}๐_{m+\frac{1}{2},n+\frac{1}{2}}^{j+\frac{1}{2}}y.`$ |
$`Proof`$ : Put $`j_2=1/2,j=j_1+1/2`$ in the relation (3.13), then
$`\sqrt{j_1+k_1+1}๐_{k_1+\frac{1}{2},m}^{j_1+\frac{1}{2}}+\sqrt{j_1k_1+1}(\delta _{k_2,\frac{1}{2}}2k_1h\delta _{k_2,\frac{1}{2}})๐_{k_1\frac{1}{2},m}^{j_1+\frac{1}{2}}`$
$`=`$ $`\sqrt{j_1+m+{\displaystyle \frac{1}{2}}}๐_{k_1,m\frac{1}{2}}^{j_1}๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}+\sqrt{j_1m+{\displaystyle \frac{1}{2}}}๐_{k_1,m+\frac{1}{2}}^{j_1}(๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}(2m+1)h๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}).`$
Replacing $`j_1+\frac{1}{2}`$ and $`k_1+\frac{1}{2}`$ with $`j`$ and $`k`$, respectively, we obtain
$`\sqrt{j+k}\delta _{k_2,\frac{1}{2}}๐_{k,m}^j+\sqrt{jk+1}(\delta _{k_2,\frac{1}{2}}(2k1)h\delta _{k_2,\frac{1}{2}})๐_{k1,m}^j`$
$`=`$ $`\sqrt{j+m}๐_{k\frac{1}{2},m\frac{1}{2}}^{j\frac{1}{2}}๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}+\sqrt{jm}๐_{k\frac{1}{2},m+\frac{1}{2}}^{j\frac{1}{2}}(๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}(2m+1)h๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}).`$
The recurrence relations (i) and (ii) are obtained by putting $`k_2=1/2`$ and $`k_2=1/2`$, respectively.
We repeat the similar computation for the relation (3.14). We put $`j_2=1/2,j=j_1+1/2`$ in (3.14), then rearrange some variables. We obtain
$`\sqrt{j+n}\{\delta _{n_2,\frac{1}{2}}+(2n1)h\delta _{n_2,\frac{1}{2}}\}๐_{m,n}^j+\sqrt{jn+1}\delta _{n_2,\frac{1}{2}}๐_{m,n1}^j`$
$`=`$ $`\sqrt{j+m}๐_{m\frac{1}{2},n\frac{1}{2}}^{j\frac{1}{2}}\{๐_{\frac{1}{2},n_2}^{\frac{1}{2}}+(2m1)h๐_{\frac{1}{2},n_2}^{\frac{1}{2}}\}+\sqrt{jm}๐_{m+\frac{1}{2},n\frac{1}{2}}^{j\frac{1}{2}}๐_{\frac{1}{2},n_2}^{\frac{1}{2}}.`$
The recurrence relations (iii) and (iv) correspond to the cases of $`n_2=1/2`$ and $`n_2=1/2`$, respectively.
The recurrence relations (v) - (viii) correspond to $`j_2=1/2,j=j_11/2`$. In this case, the relation (3.13) yields after rearrangement of variables
$`\sqrt{jk+1}\delta _{k_2,\frac{1}{2}}๐_{k,m}^j\sqrt{j+k}(\delta _{k_2,\frac{1}{2}}(2k1)h\delta _{k_2,\frac{1}{2}})๐_{k1,m}^j`$
$`=`$ $`\sqrt{jm+1}๐_{k\frac{1}{2},m\frac{1}{2}}^{j+\frac{1}{2}}๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}\sqrt{j+m+1}๐_{k\frac{1}{2},m+\frac{1}{2}}^{j+\frac{1}{2}}(๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}(2m+1)h๐_{k_2,\frac{1}{2}}^{\frac{1}{2}}).`$
Putting $`k_2=1/2`$ and $`1/2`$, we obtain the relations (v) and (vi), respectively. The relation (3.14) yields
$`\sqrt{jn+1}(\delta _{n_2,\frac{1}{2}}+(2n1)h\delta _{n_2,\frac{1}{2}})๐_{m,n}^j\sqrt{j+n}\delta _{n_2,\frac{1}{2}}๐_{m,n1}^j`$
$`=`$ $`\sqrt{jm+1}๐_{m\frac{1}{2},n\frac{1}{2}}^{j+\frac{1}{2}}(๐_{\frac{1}{2},n_2}^{\frac{1}{2}}+(2m1)h๐_{\frac{1}{2},n_2}^{\frac{1}{2}})\sqrt{j+m+1}๐_{m+\frac{1}{2},n\frac{1}{2}}^{j+\frac{1}{2}}๐_{\frac{1}{2},n_2}^{\frac{1}{2}}.`$
The recurrence relations (vii) and (viii) are obtained as the cases of $`n_2=1/2`$ and $`n_2=1/2`$, respectively. $`\mathrm{}`$
It is possible to obtain the explicit form of $`D`$-functions for some special cases such as $`๐_{m^{},j}^j,๐_{j,m}^j`$, by solving these recurrence relations. However, it seems to be difficult to derive formulae for $`๐_{m^{},m}^j`$ for any values of $`j,m^{}`$ and $`m`$. We will solve this problem by using the tensor operator approach in ยงV.
The orthogonality-like relations for $`๐_{m^{},m}^j`$ can be obtained from the relations (3.13) and (3.14).
###### Proposition 3.6
The $`D`$-functions for $`SL_h(2)`$ $`๐_{m^{},m}^j`$ satisfy the orthogonality-like relations which are reduced to the ortogonality relations of $`SL(2)`$ $`D`$-functions in the limit of $`h=0`$.
$`{\displaystyle \underset{m_1,m_2}{}}(1)^{k_1m_1}(F^{j,j})_{m_1,m_2}^{m1,m1}๐_{k_1,m_1}^j๐_{k_2,m_2}^j=(F^{j,j})_{k_1,k_2}^{k_1,k_1},`$ (3.21)
$`{\displaystyle \underset{k_1,k_2}{}}(1)^{m_1k_1}((F^1)^{j,j})_{k_1,k_1}^{k_1,k_2}๐_{k_1,m_1}^j๐_{k_2,m_2}^j=((F^1)^{j,j})_{m_1,m_1}^{m_1,m_2}.`$ (3.22)
$`Proof`$ : Consider the cases of $`j=0,j_1=j_2`$ in the relations (3.13) and (3.14). Writing $`j_1=j_2=j`$, they yield
$`{\displaystyle \underset{m_1,m_2}{}}\mathrm{\Omega }_{m_1,m_2,0}^{j,j,\mathrm{\hspace{0.33em}0}}๐_{k_1,m_1}^j๐_{k_2,m_2}^j=\mathrm{\Omega }_{k_1,k_2,0}^{j,j,0},`$
$`{\displaystyle \underset{k_1,k_2}{}}\mathrm{}_{k_1,k_2,0}^{j,j,\mathrm{\hspace{0.33em}0}}๐_{k_1,m_1}^j๐_{k_2,m_2}^j=\mathrm{}_{m_1,m_2,0}^{j,j,\mathrm{\hspace{0.33em}0}}.`$
The CGC are given by
$$\mathrm{\Omega }_{m_1,m_2,0}^{j,j,\mathrm{\hspace{0.33em}0}}=\underset{s}{}C_{s,s,0}^{j,j,0}(F^{j,j})_{m_1,m_2}^{s,s},\mathrm{}_{m_1,m_2,0}^{j,j,\mathrm{\hspace{0.33em}0}}=\underset{s}{}C_{s,s,0}^{j,j,0}((F^1)^{j,j})_{s,s}^{m_1,m_2},$$
and
$$(F^{j,j})_{m_1,m_2}^{s,s}=\delta _{s,m_1}jm_2\left|e^{s\sigma }\right|js,((F^1)^{j,j})_{s,s}^{m_1,m_2}=\delta _{s,m_1}js\left|e^{m_1\sigma }\right|jm_2.$$
Then the proof of Proposition 3.6 is straightforward. $`\mathrm{}`$
## IV. Review of $`๐บ๐ณ\mathbf{(}\mathrm{๐}\mathbf{)}`$ Representation Functions
This section is devoted to a review of the $`D`$-functions for Lie group $`SL(2).`$ Especially, we focus on tensor operator properties and the relationship to Jacobi polynomials. We write the $`D`$-functions for $`SL(2)`$ in terms of boson operators for the viewpoint of tensor operators.
Let $`a_i^j,\overline{a}_i^j,i,j\{1,2\}`$ be four copies of a boson operator commuting one another, $`i.e.`$,
$$[\overline{a}_i^j,a_k^{\mathrm{}}]=\delta _{i,k}\delta ^{j,\mathrm{}},[a_i^j,a_k^{\mathrm{}}]=[\overline{a}_i^j,\overline{a}_k^{\mathrm{}}]=0.$$
(4.1)
It is known that the Lie algebra $`gl(2)gl(2)`$ is realized by these boson operators. The left (lower) generators are defined by
$$E_{ij}=a_i^1\overline{a}_j^1+a_i^2\overline{a}_j^2,$$
(4.2)
the right (upper) generators are defined by
$$E^{ij}=a_1^i\overline{a}_1^j+a_2^i\overline{a}_2^j.$$
(4.3)
Then both left and right generators satisfy the $`gl(2)`$ commutation relations and furthermore $`[E_{ij},E^{k,\mathrm{}}]=0.`$ Each $`gl(2)`$ has decomposition $`gl(2)=sl(2)u(1)`$. The left and right $`sl(2)`$ are generated by
$$J_+=E_{21},J_{}=E_{12},J_0=E_{22}E_{11},$$
(4.4)
and
$$K_+=E^{12},K_{}=E^{21},K_0=E^{11}E^{22},$$
(4.5)
respectively, and $`u(1)`$ sectors by $`Z_L=E_{11}E_{22}`$ and $`Z_R=E^{11}+E^{22}`$. This choice of generators may be different from the usual one (see for example Ref. ยง4.4). However this is a suitable choice for twisting discussed in the next section. Note also that, in this realization, $`Z_L=Z_R.`$ Therefore, strictly speaking, this realization is not the direct sum of two copies of $`gl(2)`$.
The $`D`$-functions for Lie group $`GL(2)`$ can be given in terms of $`a_i^j`$
$$๐_{m^{},m}^{(0)j}=\{(j+m^{})!(jm^{})!(j+m)!(jm)!\}^{1/2}\underset{K,L,M,N}{}\frac{(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^N}{K!L!M!N!},$$
(4.6)
where the sum over $`K,L,M`$ and $`N`$ runs nonnegative integers provided that
$$\begin{array}{cc}K+L=j+m,\hfill & M+N=jm,\hfill \\ K+M=j+m^{},\hfill & L+N=jm.^{}\hfill \end{array}$$
(4.7)
We obtain $`SL(2)`$ $`D`$-functions by imposing $`a_1^1a_2^2a_2^1a_1^2=1`$.
It is not difficult to see that $`D`$-functions (4.6) form the irreducible tensor operators for both left and right $`gl(2)`$, $`i.e.`$,
$`[J_\pm ,๐_{m^{},m}^{(0)j}]=\sqrt{(j\pm m^{})(jm^{}+1)}๐_{m^{}1,m}^{(0)j},`$
$`[J_0,๐_{m^{},m}^{(0)j}]=2m^{}๐_{m^{},m}^{(0)j},[Z_L,๐_{m^{},m}^{(0)j}]=2j๐_{m^{},m}^{(0)j},`$ (4.8)
and
$`[K_\pm ,๐_{m^{},m}^{(0)j}]=\sqrt{(jm)(j\pm m+1)}๐_{m^{},m\pm 1}^{(0)j},`$
$`[K_0,๐_{m^{},m}^{(0)j}]=2m๐_{m^{},m}^{(0)j},[Z_R,๐_{m^{},m}^{(0)j}]=2j๐_{m^{},m}^{(0)j}.`$ (4.9)
It is well known that the $`D`$-functions for $`SL(2)`$ can be expressed in terms of the Jacobi polynomials. The Jacobi polynomials are defined by
$$P_n^{(\alpha ,\beta )}(z)=\underset{r0}{}\frac{(n)_r(\alpha +\beta +n+1)_r}{(1)_r(\alpha +1)_r}z^r,$$
(4.10)
where $`(\alpha )_r`$ stands for the sifted factorial
$$(\alpha )_r=\alpha (\alpha +1)\mathrm{}(\alpha +r1).$$
For the case of $`SL(2)`$, we have the relation $`a_1^1a_2^2=1+a_2^1a_1^2`$. Using this, the $`D`$-functions are expressed for $`m^{}+m0,m^{}m`$
$$๐_{m^{},m}^{(0)j}=\left\{\left(\begin{array}{c}j+m^{}\\ m^{}m\end{array}\right)\left(\begin{array}{c}jm\\ m^{}m\end{array}\right)\right\}^{1/2}(a_1^1)^{m^{}+m}(a_1^2)^{m^{}m}P_{jm^{}}^{(m^{}m,m^{}+m)}(z),$$
(4.11)
where $`za_2^1a_1^2`$. We have the similar relations for other cases.
## V. Representation Functions for $`๐บ๐ณ_๐\mathbf{(}\mathrm{๐}\mathbf{)}`$
### A. Explicit Formulae for $`D`$-Functions
We saw, in the previous section, that the $`D`$-functions for $`GL(2)`$ form the irreducible tensor operators of both left and right $`gl(2)`$. This fact leads us to the expectation that the $`D`$-functions for $`GL_h(2)`$ also form the irreducible tensor operators of left and right $`๐ฐ_h(gl(2))`$. It is known that the tensor operators for $`๐ฐ_h(gl(2))`$ can be obtained from the ones for $`gl(2)`$ by twisting . Therefore, we may obtain the $`D`$-functions for $`GL_h(2)`$ from the one for $`GL(2)`$ by twisting twice. The irreducible tensor operators for $`๐ฐ_h(gl(2))`$ are defined by replacing the comutator on the left hand side of (4.8) and (4.9) with the adjoint action. Let $`๐ญ`$ be a any tensor operator for $`๐ฐ_h(gl(2))`$ and $`X๐ฐ_h(gl(2))`$, then the adjoint action of $`X`$ on $`๐ญ`$ is defined by
$$\mathrm{ad}X(๐ญ)=m(idS)(\mathrm{\Delta }(X)(๐ญ1)).$$
(5.1)
The tensor operators $`๐ญ`$ for $`๐ฐ_h(gl(2))`$ and the tensor operators $`๐ญ^{(0)}`$ for $`gl(2)`$ are related via the twist element $``$ by the relation (see also Ref. )
$$๐ญ=m(idS)((๐ญ^{(0)}1)^1).$$
(5.2)
Note that $`gl(2)`$ and $`๐ฐ_h(gl(2))`$ have the same commutation relations so that the realization (4.2) and (4.3) is the realization of $`๐ฐ_h(gl(2))`$ as well. We consider the tensor operators under this realization of $`๐ฐ_h(gl(2)).`$
Let us first consider the simplest case : $`j=1/2.`$ What we obtain in this case from (4.6), (4.8) and (4.9) is that the pairs $`(a_1^1,a_2^1),(a_1^2,a_2^2)`$ are spinors of the left $`gl(2)`$ and the pairs $`(a_1^1,a_1^2),(a_2^1,a_2^2)`$ are spinors of the right $`gl(2).`$ Namely, each boson operator $`a_i^j`$ is a component of spinor for both left and right $`gl(2)`$. This fact tells us that, by twisting via the elements
$$_L=\mathrm{exp}(\frac{1}{2}J_0\sigma _L)_,_R=\mathrm{exp}(\frac{1}{2}K_0\sigma _R)_.$$
(5.3)
with $`\sigma _L=\mathrm{ln}(12hJ_+),\sigma _R=\mathrm{ln}(12hK_+)`$, we obtain a element of spinor for both left and right $`๐ฐ_h(sl(2)).`$ To this end, it is convenient to rewrite the relation (5.2) into different form. Let us write the twist element and its inverse as
$$=\underset{a}{}f^af_a,^1=\underset{a}{}g^ag_a,$$
then
$$\mu =\underset{a}{}f^aS_0(f_a),\mu ^1=\underset{a}{}S_0(g^a)g_a.$$
Noting the identity
$$g^b\mu S_0(g_b)=g^bf^aS_0(g_bf_a)=m(idS_0)(^1)=1,$$
the relation (5.2) yields
$$๐ญ=f^a๐ญ^{(0)}g^bS(f_ag_b)=f^a๐ญ^{(0)}\mu S_0(f_ag_b)\mu ^1=f^a๐ญ^{(0)}S_0(f_a)\mu ^1.$$
(5.4)
From (5.4), the twisting by $`_L`$ reads
$$\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\left(\frac{1}{2}\right)^nJ_0^na_i^jS_0(\sigma _L)\mu ^1=a_i^j\underset{k=0}{\overset{\mathrm{}}{}}\frac{(1)^{ik}}{k!}\left(\frac{1}{2}\right)^kS(\sigma _L)^k=a_i^j\mathrm{exp}\{(1)^i\sigma _L/2\}.$$
We used the fact $`S(\sigma _L)=\sigma _L`$ in the last equality. To twist the above obtained result by $`_R`$, we can repeat the similar computation. Then we have the doubly twisted boson operators
$$a_i^j\mathrm{exp}\{(1)^i\sigma _L/2+(1)^{j+1}\sigma _R/2\}.$$
(5.5)
The commutation relations of the twisted boson operators (5.5) are obtained by straightforward computation and it shows that the twisted boson operators give a realization of generators of $`GL_h(2)`$.
###### Proposition 5.1
Let
$$\begin{array}{cc}x=a_1^1e^{(\sigma _L+\sigma _R)/2},\hfill & u=a_1^2e^{(\sigma _L+\sigma _R)/2},\hfill \\ v=a_2^1e^{(\sigma _L+\sigma _R)/2},\hfill & y=a_2^2e^{(\sigma _L\sigma _R)/2},\hfill \end{array}$$
(5.6)
then, $`x,u,v`$ and $`y`$ satisfy the commutation realtions of the generators of $`GL_h(2)`$ (2.1). In this realization, the central element $`D`$ is given
$$Dxyuvhxv=a_1^1a_2^2a_2^1a_1^2.$$
(5.7)
Note that the central element $`D`$ remains undeformed in this realization.
$`Proof`$ : One can verify the commutation relations directly. We here give some useful commutation relations for the verification. The commutation relations between $`\sigma _L,\sigma _R`$ and boson operators.
| $`[\sigma _L,a_1^1]=2he^{\sigma _L}a_2^1`$, | $`[\sigma _L,a_1^2]=2he^{\sigma _L}a_2^2`$, |
| --- | --- |
| $`[\sigma _R,a_1^2]=2he^{\sigma _R}a_1^1`$, | $`[\sigma _R,a_2^2]=2he^{\sigma _R}a_2^1`$. |
These are easily verified by using the power series expansion of $`\sigma _L,\sigma _R`$ : $`\sigma _L={\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(2hJ_+)^n}{n}}_.`$ These relations can be used to prove the following commutation relations which hold for any real $`k`$
| $`[e^{k\sigma _L},a_1^1]=2hke^{(k+1)\sigma _L}a_2^1,`$ | $`[e^{k\sigma _L},a_1^2]=2hke^{(k+1)\sigma _L}a_2^2,`$ |
| --- | --- |
| $`[e^{k\sigma _R},a_1^2]=2hke^{(k+1)\sigma _R}a_1^1,`$ | $`[e^{k\sigma _R},a_2^2]=2hke^{(k+1)\sigma _R}a_2^1.`$ |
(5.8)
$`\mathrm{}`$
Next let us consider the twisting of $`๐_{m^{},m}^{(0)j}`$ for any values of $`j`$ by the twist elements $`_L,_R.`$ We denote the doubly twisted $`๐_{m^{},m}^{(0)j}`$ by $`๐_{m^{},m}^j`$, since it will be shown later that this $`๐_{m^{},m}^j`$ gives the $`D`$-functions for $`GL_h(2)`$. The computation is almost same as the case of spinors. What we need to compute is the twisting of $`(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^N`$ in the expression (4.6). The twisting by $`_L`$ reads
$`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{1}{2}}\right)^nJ_0^n(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^NS_0^n(\sigma _L)\mu ^1`$
$`=`$ $`(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^N{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{k!}}\left({\displaystyle \frac{1}{2}}\right)^k(K+LM+N)^k\mu S_0^k(\sigma _L)\mu ^1`$
$`=`$ $`(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^N\mathrm{exp}\{(KL+MN)\sigma _L/2\}.`$
Further twisting by $`_R`$ gives
$$(a_1^1)^K(a_2^1)^L(a_1^2)^M(a_2^2)^N\mathrm{exp}\{(KL+MN)\sigma _L/2+(K+LMN)\sigma _R/2\}.$$
(5.9)
Because of the condition (4.7), we have $`KL+MN=2m^{}`$ and $`K+LMN=2m.`$ Thus the exponential factor appeared in (5.9) is factored out the sum over $`K,L,M`$ and $`N`$. Therefore we have proved the following proposition.
###### Proposition 5.2
In the realization (4.2), (4.3), the irreducible tensor operators of both left and right $`๐ฐ_h(gl(2))`$ are given by
$$๐_{m^{},m}^j=๐_{m^{},m}^{(0)j}e^{m^{}\sigma _L+m\sigma _R}.$$
(5.10)
One can write $`๐_{m^{},m}^j`$ of Proposition 5.10 in terms of the generators of $`GL_h(2)`$ by making use of Proposition 5.1. For real $`A,B`$,
$`(a_1^1)^Ke^{(A\sigma _L+B\sigma _R)/2}=(a_1^1)^{K1}xe^{(A+1)\sigma _L/2+(B1)\sigma _R/2}`$
$`=`$ $`(a_1^1)^{K1}e^{(A+1)\sigma _L/2+(B1)\sigma _R/2}\{e^{(A+1)\sigma _L/2}xe^{(A+1)\sigma _L/2}\}.`$
The expression $`\{\mathrm{}\}`$ in the last line can be calculated by using (5.8) and gives $`xh(A+1)v`$. Thus we have obtained
$`(a_1^1)^Ke^{(A\sigma _L+B\sigma _R)/2}=e^{(A+K)\sigma _L/2+(BK)\sigma _R/2}`$
$`\times (xh(A+K)v)(xh(A+K1)v)\mathrm{}(xh(A+1)v).`$ (5.11)
Similar computation gives three other identities
$`(a_2^1)^Le^{(A\sigma _L+B\sigma _R)/2}=e^{(AL)\sigma _L/2+(BL)\sigma _R/2}v^L,`$
$`(a_1^2)^Me^{(A\sigma _L+B\sigma _R)/2}=e^{(A+M)\sigma _L/2+(B+M)\sigma _R/2}`$
$`\times (uh(B+M)xh(A+M)y+h^2(A+M)(B+M)v)`$
$`\times (uh(B+M1)xh(A+M1)y+h^2(A+M1)(B+M1)v)`$
$`\times \mathrm{}\times (uh(B+1)xh(A+1)y+h^2(A+1)(B+1)v),`$ (5.12)
$`(a_2^2)^Ne^{(A\sigma _L+B\sigma _R)/2}=e^{(AN)\sigma _L/2+(B+N)\sigma _R/2}`$
$`\times (yh(B+N)v)(hh(B+N1)v)\mathrm{}(yh(B+1)v).`$
The boson operators $`a_i^j`$ commute one another so that the order of $`a_i^j`$โs in $`๐_{m^{},m}^{(0)j}`$ is irrelevant. Therefore we can have some different expressions of $`๐_{m^{},m}^j`$ according to the choice of the order of boson operators. We here give two of them and shall show that they are the representation functions of $`GL_h(2)`$.
###### Proposition 5.3
The $`D`$-functions for $`GL_h(2)`$ are given by
$$๐_{m^{},m}^j=\{(j+m^{})!(jm^{})!(j+m)!(jm)!\}^{1/2}\underset{K,L,M,L}{}\frac{X_Kv^LU_{K,L,M}Y_{K,L,M,N}}{K!M!L!N!},$$
(5.13)
where $`X_K,U_{K,L,M}`$ and $`Y_{K,L,M,N}`$ are defined by
$`X_K`$ $`=`$ $`x(x+hv)\mathrm{}(x+h(K1)v),`$
$`U_{K,L,M}`$ $`=`$ $`(uh(K+L)x+h(KL)yh^2(K^2L^2)v)`$
$`\times `$ $`(uh(K+L1)x+h(KL+1)yh^2(K^2(L1)^2)v)`$
$`\times `$ $`\mathrm{}`$
$`\times `$ $`(uh(K+LM+1)x+h(KL+M1)y`$
$`h^2(K^2(LM+1)^2)v)`$
$`Y_{K,L,M,N}`$ $`=`$ $`(yh(K+LM)v)(yh(K+LM1)v)`$
$`\times `$ $`\mathrm{}(yh(K+LMN+1)v).`$
The $`D`$-functions have another expression which is
$$๐_{m^{},m}^j=\{(j+m^{})!(jm^{})!(j+m)!(jm)!\}^{1/2}\underset{K,L,M,L}{}\frac{U_MX_{K,M}Y_{K,M,N}v^L}{K!M!L!N!},$$
(5.14)
where $`U_M,X_{K,M},Y_{K,M,N}`$ are defined by
$`U_M`$ $`=`$ $`u(u+h(x+y)+h^2v)\mathrm{}(u+h(M1)(x+y)+h^2(M1)^2v),`$
$`X_{K,M}`$ $`=`$ $`(x+hMv)(x+h(M+1)v)\mathrm{}(x+h(K+M1)v),`$
$`Y_{K,M,N}`$ $`=`$ $`(yh(KM)v)(yh(KM1)v)\mathrm{}(yh(KMN+1)v)v^L.`$
The sum over $`K,L,M`$ and $`N`$ runs nonnegative integers under the condition (4.7).
$`Remark`$ : We obtain the $`D`$-functions for $`SL_h(2)`$ by putting $`D=xyuvhxv=1.`$
$`Proof`$ : These expressions are obtained by using (5.11) and (5.12). The expression (5.13) corresponds to the boson ordering $`(a_1^1)^k(a_2^1)^L(a_1^2)^M(a_2^2)^N`$, while the expression (5.14) corresponds to $`(a_1^2)^M(a_1^1)^K(a_2^2)^N(a_2^1)^L.`$
To show that $`๐_{m^{},m}^j`$ are the representation functions of $`GL_h(2)`$, we must verify (2.8). It is obvious that $`๐_{m^{},m}^jGL_h(2)`$ and the counit of $`๐_{m^{},m}^j`$ is easily verified by using $`ฯต(x)=ฯต(y)=1,ฯต(u)=ฯต(v)=0`$. However, it seems to be difficult to verify the coproduct of $`๐_{m^{},m}^j`$ by straightforward computation. Instead of verifying the coproduct, we show that $`๐_{m^{},m}^j`$ satisfy the recurrence relations of Proposition 3.5. Note that the recurrence relations of Proposition 3.5 are for $`SL_h(2)`$. The Jordanian deformation of the Lie algebra $`gl(2)`$ considered in this paper is the direct sum of the deformed $`sl(2)`$ and undeformed $`u(1)`$ : $`๐ฐ_h(gl(2))=๐ฐ_h(sl(2))u(1).`$ This implies that the CGC for $`๐ฐ_h(sl(2))`$ also give the CGC for $`๐ฐ_h(gl(2))`$. Therefore the $`D`$-functions for $`GL_h(2)`$ also satisfy the recurrence relations of Proposition 3.5.
As an example, we show that the $`๐_{m^{},m}^j`$ give the solutions to the recurrence relation (ii). We substitute the expression (5.14) of the $`D`$-functions into the first term of the RHS of the relation (ii), then replace the dummy index $`L`$ with $`L1`$. It follows that
$$\sqrt{j+m}๐_{k\frac{1}{2},m\frac{1}{2}}^{j\frac{1}{2}}v=\{(j+m)!(jm)!(j+k1)!(jk)!\}^{1/2}\underset{K,L,M,N}{}L\frac{U_MX_{K,M}Y_{K,M,N}v^L}{K!M!L!N!}_,$$
where the indices $`K,L,M`$ and $`N`$ satifsy the condition
$$K+L=j+m,M+N=jm,K+M=j+k1,L+N=jk+1.$$
(5.15)
For the second term of the RHS of the relation (ii), we use the expression (5.13). Replacing the index $`N`$ with $`N1`$, we obtain
$`\sqrt{jm}๐_{k\frac{1}{2},m+\frac{1}{2}}^{j\frac{1}{2}}(u(2m+1)hv)`$
$`=\{(j+m)!(jm)!(j+k1)!(jk)!\}^{1/2}{\displaystyle \underset{K,L,M,N}{}}N{\displaystyle \frac{X_Kv^LU_{K,L,M}Y_{K,L,M,N}}{K!M!L!N!}}_,`$
where the indices $`K,L,M`$ and $`N`$ also satisfy the condition (5.15). Since the expressions (5.13) and (5.14) are the different expressions of the same $`D`$-functions, it holds that $`U_MX_{K,M}Y_{K,M,N}v^L=X_Kv^LU_{K,L,M}Y_{K,L,M,N}.`$ Therefore the RHS of (ii) reads
$`\{(j+m)!(jm)!(j+k1)!(jk)!\}^{1/2}{\displaystyle \underset{K,L,M,N}{}}(L+N){\displaystyle \frac{X_Kv^LU_{K,L,M}Y_{K,L,M,N}}{K!M!L!N!}}`$
$`=`$ $`\sqrt{jk+1}๐_{k1,m}^j.`$
The four-term recurrence relation (i) is reduced to a three-term relation, by eliminating $`๐_{k1,m}^j`$ from (i) and (ii). This recurrence relation is easily solved by using the another expression of $`๐_{k,m}^j`$ corresponding to another ordering of boson operators. The suitable expressions for solving it are the ones obtained from the ordering $`(a_1^2)^M(a_2^2)^N(a_2^1)^L(a_1^1)^K`$ and $`(a_1^1)^K(a_2^2)^N(a_2^1)^L(a_1^2)^M.`$ In this way, we can verify the $`๐_{m^{},m}^j`$ obtained in this Proposition solve all the recurrence relations given in Proposition 3.5. $`\mathrm{}`$
Both expression of (5.13) and (5.14), of course, give the generators of $`GL_h(2)`$ for $`j=1/2`$ which reflects Proposition 3.20. The $`D`$-functions for $`j=1`$ reads
$$๐^1=\left(\begin{array}{ccc}x^2+hxv& \sqrt{2}(ux+huv)& u^2+h(ux+uy+huv)\\ \sqrt{2}xv& D+2uv& \sqrt{2}(uy+huv)\\ v^2& \sqrt{2}yv& y^2+hyv\end{array}\right)_.$$
(5.16)
For $`SL_h(2),i.e.`$ putting $`D=1`$, this coincides with the one obtained by using $`h`$-symplecton or quantum $`h`$-plane . Chakrabarti and Quesne obtained the $`๐^1`$ for two-parametric Jordanian deformation of $`GL(2)`$ in the coloured representation through a contraction technique to the $`D`$-functions for standard $`(q,\lambda )`$-deformation of $`GL(2)`$ . To compare the present $`๐^1`$ with the one given in Ref., put $`\alpha =0,z=1`$ in Eqs.(4.20) and (4.21) of Ref.. Then we see that the $`D`$-functions for $`j=1`$ of Ref. are different form (5.16). This difference stems from the different choice of the basis of $`๐ฐ_h(sl(2))`$. In Ref., the basis introduced by Ohn is used, that is, the commutation relations of the generators of $`๐ฐ_h(sl(2))`$ are not same as those of $`sl(2)`$. While the basis of this paper satisfy the same commutation relations as $`sl(2)`$. This results the different CGC for the same algebra so that the recurrence relations for the $`D`$-functions have the different form. The CGC for Ohnโs basis are found in Ref.. Repeating the same procedure as ยงIII.B, we obtain another form of recurrence relations. It may be easy to verify that the $`๐^1`$ of Ref. solves these recurrence relations.
### B. $`๐บ๐ณ_๐\mathbf{(}\mathrm{๐}\mathbf{)}`$ $`D`$-Functions and Jacobi Polynomials
The purpose of this subsection is to show that the $`D`$-functions for $`SL_h(2)`$ can be expressed in terms of Jacobi polynomials. To this end, we return to the boson realization of $`D`$-functions (Proposition 5.10) and use the fact that the $`D`$-functions for Lie group $`SL(2)`$ are written in terms of Jacobi polynomials. Recall the following two facts : (1) the central element $`D`$ of $`GL_h(2)`$ is not deformed in the boson realization, Eq.(5.7), (2) Jacobi polynomials appeared in the $`D`$-functions for $`SL(2)`$ are power series in the variable $`z=a_2^1a_1^2.`$ We write the $`D`$-functions $`๐_{m^{},m}^{(0)j}`$ for $`SL(2)`$ appeared in (5.10) in terms of Jacobi polynomials then use the easily proved relation $`(a_2^1a_1^2)^r=(uv)^r`$ in order to replace the variable $`z=a_2^1a_1^2`$ with the $`h`$-deformed one $`z=uv`$. Let us consider, as an example, the case of $`m^{}+m0,m^{}m`$. The $`๐_{m^{},m}^{(0)j}`$ are given by (4.11). We rearrange the order of $`a_1^1,a_1^2`$ and $`P_{jm^{}}^{(m^{}m,m^{}+m)}(z)`$ to be $`P_{jm^{}}^{(m^{}m,m^{}+m)}(z)(a_1^2)^{m^{}m}(a_1^1)^{m^{}+m}.`$ Using the relations (5.11) and (5.12), we see that
$`(a_1^2)^{m^{}m}(a_1^1)^{m^{}+m}e^{m^{}\sigma _L+m\sigma _R}`$
$`=`$ $`u(u+h(x+y)+h^2v)\mathrm{}(u+h(m^{}m1)(x+y)+h^2(m^{}m1)^2v)`$
$`\times `$ $`(x+h(m^{}m)v)(x+h(m^{}m1)v)\mathrm{}(x+h(2m^{}1)v).`$
This completes the expression of $`D`$-functions in terms of Jacobi polynomials.
Repeating this process for other cases, we can prove the next proposition.
###### Proposition 5.4
The $`D`$-functions for $`SL_h(2)`$ are written in terms of Jacobi polynomials as follows:
(i) $`m^{}+m0,m^{}m`$
$`๐_{m^{},m}^j`$ $`=`$ $`N_+P_{jm^{}}^{(m^{}m,m^{}+m)}(z)`$
$`\times `$ $`u(u+h(x+y)+h^2v)\mathrm{}(u+h(m^{}m1)(x+y)+h^2(m^{}m1)^2v)`$
$`\times `$ $`(x+h(m^{}m)v)(x+h(m^{}m1)v)\mathrm{}(x+h(2m^{}1)v).`$
(ii) $`m^{}+m0,m^{}m`$
$$๐_{m^{},m}^j=N_{}P_{jm}^{(m^{}+m,m^{}+m)}(z)x(x+hv)\mathrm{}(x+h(m^{}+m1)v)v^{m^{}+m}$$
(iii) $`m^{}+m0,m^{}m`$
$`๐_{m^{},m}^j`$ $`=`$ $`N_+P_{j+m}^{(m^{}m,m^{}m)}(z)`$
$`\times `$ $`u(u+h(x+y)+h^2v)\mathrm{}(u+h(m^{}m1)(x+y)+h^2(m^{}m1)^2v)`$
$`\times `$ $`(yh(mm^{})v)(yh(mm^{}1)v)\mathrm{}(yh(2m+1)v).`$
(iv) $`m^{}+m0,m^{}m`$
$`๐_{m^{},m}^j`$ $`=`$ $`N_{}P_{j+m^{}}^{(m^{}+m,m^{}m)}(z)`$
$`\times `$ $`v^{m^{}+m}(yh(mm^{})v)(yh(mm^{}1)v)\mathrm{}(yh(2m+1)v).`$
The variable $`z`$ is defined by $`z=uv`$ and the factors $`N_+,N_{}`$ by
$$N_+=\left\{\left(\begin{array}{c}j+m^{}\\ m^{}m\end{array}\right)\left(\begin{array}{c}jm\\ m^{}m\end{array}\right)\right\}_,^{1/2}N_{}=\left\{\left(\begin{array}{c}jm^{}\\ mm^{}\end{array}\right)\left(\begin{array}{c}j+m\\ mm^{}\end{array}\right)\right\}_.^{1/2}$$
$`Remark`$ : The Jacobi polynomials are to the left of the generators of $`SL_h(2)`$. To move $`P_n^{(\alpha ,\beta )}(z)`$ to the right, the relation
$$(uv)^r\mathrm{exp}(m^{}\sigma _L+m\sigma _R)=\mathrm{exp}(m^{}\sigma _L+m\sigma _R)\{uv2h(m^{}yv+mxv)4h^2mm^{}v^2\}^r,$$
is used and we see that the Jacobi polynomials are changed to the power series in $`\zeta _{m^{},m}=(u+2h(m^{}ymx)4h^2mm^{})v`$, but the rests of the formulae remain unchanged.
## VI. Boson Realization of $`๐ฎ๐ณ_{๐\mathbf{,}๐}\mathbf{(}\mathrm{๐}\mathbf{)}`$
It is natural to generalize the results in the previous section to the two-parametric Jordanian deformation of $`GL(2)`$ , since the twist element which generates the two-parametric Jordanian quantum algebra $`๐ฐ_{h,g}(gl(2))`$ is known . Unfortunately, the method in the previous sections leads us to quite complex calculation. As the first step to obtain the $`D`$-functions for two-parametric Jordanian quantum group $`GL_{h,g}(2)`$, we here give the boson realization of the generators of $`GL_{h,g}(2)`$.
The left and right twist elements are given by
$`_L=\mathrm{exp}\left({\displaystyle \frac{g}{2h}}\sigma _LZ_L\right)\mathrm{exp}\left({\displaystyle \frac{1}{2}}J_0\sigma _L\right),`$
$`_R=\mathrm{exp}\left({\displaystyle \frac{g}{2h}}\sigma _RZ_R\right)\mathrm{exp}\left({\displaystyle \frac{1}{2}}K_0\sigma _R\right),`$
respectively. We can see that the $`GL_{h,g}(2)`$ is reduced to $`GL_h(2)`$ when $`g=0`$. Repeating the same procedure as (5.5), we obtain the twisted boson operators. We can rewrite the twisted boson operator in terms of the generators $`GL_h(2)`$. The next proposition can be regarded as a realization of $`GL_{h,g}(2)`$ by generators of $`GL_h(2)`$ and $`Z_L,Z_R`$ as well.
###### Proposition 6.1
Let
$$\begin{array}{cc}a=xgvZ_L,\hfill & b=ugxZ_RgyZ_L+g^2vZ_LZ_R,\hfill \\ c=v,\hfill & d=ygvZ_R,\hfill \end{array}$$
(6.1)
where $`x,u,v`$ and $`y`$ are given by (5.6). Then $`a,b,c`$ and $`d`$ satisfy the commutation relation of $`GL_{h,g}(2)`$.
$`Remark`$ : In this realization, the quantum determinant $`D^{}=adbc(h+g)ac`$ for $`GL_{h,g}(2)`$ and $`D`$ for $`GL_h(2)`$ coincide : $`D^{}=D=a_1^1a_2^2a_2^1a_1^2`$.
$`Proof`$ : It requires a lengthy calculation, however, the proof is straightforward. The following commutation relations are verified.
$`[a,b]=(h+g)(D^{}a^2),[a,c]=(hg)c^2,`$
$`[a,d]=(h+g)ac(hg)dc,[b,c]=(h+g)ac(hg)cd,`$ (6.2)
$`[b,d]=(hg)(D^{}d^2),[c,d]=(h+g)c^2.`$
$`\mathrm{}`$
## VII. Concluding Remarks
In this paper, the explicit formulae of the $`D`$-functions for $`SL_h(2)`$ (and $`GL_h(2)`$) have been obtained by using the tensor operator technique. We used the fact that the $`D`$-functions for Lie group $`GL(2)`$ form irreducible tensor operators of $`gl(2)gl(2)`$ in the realization (4.2), (4.3). This kind of tensor operators are called double irreducible tensor operators in the literature. The $`D`$-functions for $`GL_h(2)`$ were obtained via the construction of double irreducible tensor operators for $`๐ฐ_h(gl(2))๐ฐ_h(gl(2)).`$ Other examples of double irreducible tensor operators were considered for $`q`$-deformation and for Jordanian deformation . Quesne constructed the $`GL_h(n)\times GL_h^{}(m)`$ covariant bosonic and fermionic algebra which form the double irreducible tensor operators of $`๐ฐ_h(gl(n))๐ฐ_h^{}(gl(m))`$ using the contraction method . This may suggest, in the case of $`n=m=2`$ and $`h=h^{}`$, that the bosonic algebra of Quesne has a close relation to $`๐_{m^{},m}^{\frac{1}{2}},i.e.,`$ the generators of $`GL_h(2).`$
We also showed that the $`D`$-functions for $`SL_h(2)`$ can be expressed in terms of Jacobi polynomials. Contrary to the $`q`$-deformed case where the little $`q`$-Jacobi polynomials appear in the $`D`$-functions for $`SU_q(2)`$, the ordinary Jacobi polynomials are associated with the $`D`$-functions for $`SL_h(2)`$. It seems to be a general feature of Jordanian deformation that the ordinary orthogonal polynomials are associated with the representations. It is known that the ordinary Gauss hypergeometirc functions are associated with $`h`$-symplecton , while the $`q`$-hypergeometric functions are associated with the $`q`$-deformation of symplecton.
The extension of the results of this paper to the Jordanian deformation of $`SL(n)`$ may be possible, since the explicit expressions for the twist element are known for the Lie algebra $`sl(n)`$ . |
warning/0003/hep-ph0003022.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The search for the lightest Higgs boson is a crucial test of Supersymmetry (SUSY) that can be performed with the present and the next generation of accelerators. The prediction of a relatively light Higgs boson is common to all supersymmetric models whose couplings remain in the perturbative regime up to a very high energy scale . Finding the Higgs boson is thus one of the main goals of todays high-energy physics. Concerning the Higgs boson search, it is necessary to know the decay widths and branching ratios of the main decay channels to a high accuracy. After the detection of a scalar particle it is mandatory as a next step to measure its couplings to gauge bosons and fermions and also its self-couplings very accurately, in order to establish the Higgs mechanism and the Yukawa interactions experimentally. The determination of the trilinear Higgs-boson self-couplings might be possible at a future linear $`e^+e^{}`$ collider with high luminosity .
In this paper we concentrate on the coupling of the lightest MSSM Higgs boson to Standard model (SM) fermions. In the MSSM the mass of the lightest Higgs boson, $`m_h`$, is bounded from above by $`m_h\stackrel{<}{}\mathrm{\hspace{0.33em}135}\mathrm{GeV}`$, including radiative corrections up to two-loop order . Since the $`b`$-, the $`c`$-quark and the $`\tau `$-lepton are the heaviest particles for which the decay $`hf\overline{f}`$ is kinematically allowed, it is of particular interest to calculate their corresponding decay rates and branching ratios with high precision . We analyze these decay rates and branching ratios, taking into account the Higgs-boson propagator corrections, where at the one-loop level the full momentum dependence is kept. These corrections contain the Yukawa contributions of $`๐ช(G_Fm_t^4/M_W^2)`$, which are the dominant electroweak one-loop corrections to the Higgs-boson decay width, and the corresponding QCD corrections of $`๐ช(G_F\alpha _sm_t^4/M_W^2)`$ as well as the Yukawa corrections of $`๐ช(G_F^2m_t^6/M_W^2)`$. We also take into account the one-loop vertex corrections resulting from gluon, gluino and photon exchange together with real gluon and photon emission as given in Ref. . Only the purely weak $`๐ช(\alpha )`$ vertex corrections have been neglected. We numerically investigate the effect of the two-loop propagator contributions and the one-loop gluino-exchange vertex correction. The latter one has been mostly neglected in experimental analyses so far, but can have a large impact on the result. We show analytically that the Higgs-boson propagator correction with neglected momentum dependence can be absorbed into the tree-level coupling using the effective mixing angle from the neutral $`๐๐ซ`$-even Higgs boson sector. The result in this approximation is then compared with the full result. We also perform a comparison between our full result and the renormalization group (RG) improved effective potential calculation, see Ref. and references therein. For most parts of the MSSM parameter space we find agreement within 10% between the two approaches, although deviations up to 50% are possible for certain ranges of the parameter space.
The paper is organized as follows: in section 2 we give the calculational basis needed for the incorporation of the two-loop Higgs-propagator corrections into the decay widths. We show analytically how the propagator corrections with neglected external momentum are related to the effective mixing-angle approach. The results for the gluon, gluino and QED vertex corrections in combination with gluon and photon bremsstrahlung are reviewed. In section 3 a numerical analysis for the decay rates and a comparison of the full result and the effective mixing-angle result is performed. Special emphasis is put on the two-loop propagator correction and the one-loop gluino contribution. Section 4 contains the comparison of our full result with the RG approach. The conclusions can be found in section 5.
## 2 Calculational basis
The Higgs potential of the MSSM is given by
$`V`$ $`=`$ $`m_1^2H_1\overline{H}_1+m_2^2H_2\overline{H}_2m_{12}^2(ฯต_{ab}H_1^aH_2^b+\text{h.c.})`$ (1)
$`+{\displaystyle \frac{g^2+g^2}{8}}(H_1\overline{H}_1H_2\overline{H}_2)^2+{\displaystyle \frac{g^2}{2}}|H_1\overline{H}_2|^2,`$
where $`m_1,m_2,m_{12}`$ are soft SUSY-breaking terms, $`g,g^{}`$ are the $`SU(2)`$ and $`U(1)`$ gauge couplings, and $`ฯต_{12}=1`$. The doublet fields $`H_1`$ and $`H_2`$ are decomposed in the following way:
$`H_1`$ $`=`$ $`\left(\begin{array}{c}H_1^1\\ H_1^2\end{array}\right)=\left(\begin{array}{c}v_1+(\varphi _1^0+i\chi _1^0)/\sqrt{2}\\ \varphi _1^{}\end{array}\right),`$ (6)
$`H_2`$ $`=`$ $`\left(\begin{array}{c}H_2^1\\ H_2^2\end{array}\right)=\left(\begin{array}{c}\varphi _2^+\\ v_2+(\varphi _2^0+i\chi _2^0)/\sqrt{2}\end{array}\right).`$ (11)
Besides $`g`$ and $`g^{}`$, two independent parameters enter the potential (1): $`\mathrm{tan}\beta =v_2/v_1`$ and $`M_A^2=m_{12}^2(\mathrm{tan}\beta +\mathrm{cot}\beta )`$, where $`M_A`$ is the mass of the $`๐๐ซ`$-odd $`A`$ boson.
The $`๐๐ซ`$-even neutral mass eigenstates are obtained performing the rotation
$`\left(\begin{array}{c}H^0\\ h^0\end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}\mathrm{cos}\alpha & \mathrm{sin}\alpha \\ \mathrm{sin}\alpha & \mathrm{cos}\alpha \end{array}\right)\left(\begin{array}{c}\varphi _1^0\\ \varphi _2^0\end{array}\right)D^1(\alpha )\left(\begin{array}{c}\varphi _1^0\\ \varphi _2^0\end{array}\right)`$ (20)
with the mixing angle $`\alpha `$ related to $`\mathrm{tan}\beta `$ and $`M_A`$ by
$$\mathrm{tan}2\alpha =\mathrm{tan}2\beta \frac{M_A^2+M_Z^2}{M_A^2M_Z^2},\frac{\pi }{2}<\alpha <0.$$
(21)
At the tree level the mass matrix of the neutral $`๐๐ซ`$-even Higgs bosons in the $`\varphi _1,\varphi _2`$ basis can be expressed in terms of $`M_Z`$ and $`M_A`$ as follows:
$`M_{\mathrm{Higgs}}^{2,\mathrm{tree}}`$ $`=`$ $`\left(\begin{array}{cc}m_{\varphi _1}^2& m_{\varphi _1\varphi _2}^2\\ m_{\varphi _1\varphi _2}^2& m_{\varphi _2}^2\end{array}\right)`$ (24)
$`=`$ $`\left(\begin{array}{cc}M_A^2\mathrm{sin}^2\beta +M_Z^2\mathrm{cos}^2\beta & (M_A^2+M_Z^2)\mathrm{sin}\beta \mathrm{cos}\beta \\ (M_A^2+M_Z^2)\mathrm{sin}\beta \mathrm{cos}\beta & M_A^2\mathrm{cos}^2\beta +M_Z^2\mathrm{sin}^2\beta \end{array}\right).`$ (27)
Transforming to the eigenstate basis (20) yields
$$M_{\mathrm{Higgs}}^{2,\mathrm{tree}}\stackrel{\alpha }{}\left(\begin{array}{cc}m_H^2& 0\\ 0& m_h^2\end{array}\right)$$
(28)
with $`m_h`$ and $`m_H`$ being the tree-level masses of the neutral $`๐๐ซ`$-even Higgs bosons.
In the Feynman diagrammatic (FD) approach the higher-order corrected Higgs boson masses, denoted by $`M_h,M_H`$, are derived by finding the poles of the $`h,H`$-propagator matrix whose inverse is given by
$$\left(\mathrm{\Delta }_{\mathrm{Higgs}}\right)^1=i\left(\begin{array}{cc}q^2m_H^2+\widehat{\mathrm{\Sigma }}_H(q^2)& \widehat{\mathrm{\Sigma }}_{hH}(q^2)\\ \widehat{\mathrm{\Sigma }}_{hH}(q^2)& q^2m_h^2+\widehat{\mathrm{\Sigma }}_h(q^2)\end{array}\right),$$
(29)
where the $`\widehat{\mathrm{\Sigma }}(q^2)`$ denote the renormalized Higgs boson self-energies. For these self-energies we take the result up to two-loop order, see Sect. 2.2 below.
Our main emphasis in this paper is on the fermionic decays of the light Higgs boson, but for completeness we list the expressions for both $`h`$ and $`H`$. The amplitudes for the decays $`h,Hf\overline{f}`$ can be written as follows:
$`A(hf\overline{f})`$ $`=`$ $`\sqrt{Z_h}\left(\mathrm{\Gamma }_h+Z_{hH}\mathrm{\Gamma }_H\right),`$ (30)
$`A(Hf\overline{f})`$ $`=`$ $`\sqrt{Z_H}\left(\mathrm{\Gamma }_H+Z_{Hh}\mathrm{\Gamma }_h\right),`$ (31)
with
$`Z_{hH}`$ $`=`$ $`{\displaystyle \frac{\widehat{\mathrm{\Sigma }}_{hH}(M_h^2)}{M_h^2m_H^2+\widehat{\mathrm{\Sigma }}_H(M_h^2)}},`$ (32)
$`Z_{Hh}`$ $`=`$ $`{\displaystyle \frac{\widehat{\mathrm{\Sigma }}_{hH}(M_H^2)}{M_H^2m_h^2+\widehat{\mathrm{\Sigma }}_h(M_H^2)}},`$ (33)
involving the renormalized self-energies $`\widehat{\mathrm{\Sigma }}(q^2)`$ and 3-point vertex functions $`\mathrm{\Gamma }_h,\mathrm{\Gamma }_H`$. The wave function renormalization factors $`Z_h`$ and $`Z_H`$ are related to the finite residues of the $`h`$ and $`H`$ propagators, respectively:
$`Z_h`$ $`=`$ $`{\displaystyle \frac{1}{1+\text{Re}\widehat{\mathrm{\Sigma }}_h^{}(q^2)\text{Re}\left(\frac{\widehat{\mathrm{\Sigma }}_{hH}^2(q^2)}{q^2m_H^2+\widehat{\mathrm{\Sigma }}_H(q^2)}\right)^{}}}_{|q^2=M_h^2}`$ (34)
$`Z_H`$ $`=`$ $`{\displaystyle \frac{1}{1+\text{Re}\widehat{\mathrm{\Sigma }}_H^{}(q^2)\text{Re}\left(\frac{\widehat{\mathrm{\Sigma }}_{hH}^2(q^2)}{q^2m_h^2+\widehat{\mathrm{\Sigma }}_h(q^2)}\right)^{}}}{}_{|q^2=M_H^2}{}^{}.`$ (35)
### 2.1 The $`\alpha _{\mathrm{eff}}`$-approximation
The dominant contributions for the Higgs boson self-energies can be obtained by setting $`q^2=0`$. Approximating the renormalized Higgs boson self-energies by
$$\widehat{\mathrm{\Sigma }}(q^2)\widehat{\mathrm{\Sigma }}(0)\widehat{\mathrm{\Sigma }}$$
(36)
yields the Higgs boson masses by re-diagonalizing the dressed mass matrix
$$M_{\mathrm{Higgs}}^2=\left(\begin{array}{cc}m_H^2\widehat{\mathrm{\Sigma }}_H& \widehat{\mathrm{\Sigma }}_{hH}\\ \widehat{\mathrm{\Sigma }}_{hH}& m_h^2\widehat{\mathrm{\Sigma }}_h\end{array}\right)\stackrel{\mathrm{\Delta }\alpha }{}\left(\begin{array}{cc}M_H^2& 0\\ 0& M_h^2\end{array}\right),$$
(37)
where $`M_h`$ and $`M_H`$ are the corresponding higher-order-corrected Higgs boson masses. The rotation matrix in the transformation (37) reads:
$$D(\mathrm{\Delta }\alpha )=\left(\begin{array}{cc}\mathrm{cos}\mathrm{\Delta }\alpha & \mathrm{sin}\mathrm{\Delta }\alpha \\ \mathrm{sin}\mathrm{\Delta }\alpha & \mathrm{cos}\mathrm{\Delta }\alpha \end{array}\right).$$
(38)
The angle $`\mathrm{\Delta }\alpha `$ is related to the renormalized self-energies and masses through the eigenvector equation
$`\left(\begin{array}{cc}m_H^2\widehat{\mathrm{\Sigma }}_HM_h^2& \widehat{\mathrm{\Sigma }}_{hH}\\ \widehat{\mathrm{\Sigma }}_{hH}& m_h^2\widehat{\mathrm{\Sigma }}_hM_h^2\end{array}\right)\left(\begin{array}{c}\mathrm{sin}\mathrm{\Delta }\alpha \\ \mathrm{cos}\mathrm{\Delta }\alpha \end{array}\right)=0`$ (43)
which yields
$$\frac{\widehat{\mathrm{\Sigma }}_{hH}}{M_h^2m_H^2+\widehat{\mathrm{\Sigma }}_H}=\mathrm{tan}\mathrm{\Delta }\alpha .$$
(44)
The second eigenvector equation leads to:
$$\frac{\widehat{\mathrm{\Sigma }}_{hH}}{M_H^2m_h^2+\widehat{\mathrm{\Sigma }}_h}=\mathrm{tan}\mathrm{\Delta }\alpha .$$
(45)
With the approximation (36) one deduces
$`Z_{hH}`$ $`=`$ $`{\displaystyle \frac{\widehat{\mathrm{\Sigma }}_{hH}}{M_h^2m_H^2+\widehat{\mathrm{\Sigma }}_H}}=\mathrm{tan}\mathrm{\Delta }\alpha ,`$ (46)
$`Z_{Hh}`$ $`=`$ $`{\displaystyle \frac{\widehat{\mathrm{\Sigma }}_{hH}}{M_H^2m_h^2+\widehat{\mathrm{\Sigma }}_h}}=+\mathrm{tan}\mathrm{\Delta }\alpha ,`$ (47)
and $`Z_h`$ can be expressed as
$`Z_h`$ $`=`$ $`{\displaystyle \frac{1}{1+\left(\frac{\widehat{\mathrm{\Sigma }}_{hH}}{M_h^2m_H^2+\widehat{\mathrm{\Sigma }}_H}\right)^2}}`$ (48)
$`=`$ $`{\displaystyle \frac{1}{1+\mathrm{tan}^2\mathrm{\Delta }\alpha }}=\mathrm{cos}^2\mathrm{\Delta }\alpha .`$
Analogously one obtains
$$Z_H=\mathrm{cos}^2\mathrm{\Delta }\alpha .$$
(49)
At the tree level, the vertex functions can be written as
$`\begin{array}{c}\mathrm{\Gamma }_h=\frac{iem_f\mathrm{sin}\alpha }{2s_WM_W\mathrm{cos}\beta }=C_f^{(d)}\mathrm{sin}\alpha \hfill \\ \mathrm{\Gamma }_H=\frac{iem_f\mathrm{cos}\alpha }{2s_WM_W\mathrm{cos}\beta }=C_f^{(d)}\mathrm{cos}\alpha \hfill \end{array}\}\text{for d-type fermions}`$ (52)
$`\begin{array}{c}\mathrm{\Gamma }_h=\frac{iem_f\mathrm{cos}\alpha }{2s_WM_W\mathrm{sin}\beta }=C_f^{(u)}\mathrm{cos}\alpha ,\hfill \\ \mathrm{\Gamma }_H=\frac{iem_f\mathrm{sin}\alpha }{2s_WM_W\mathrm{sin}\beta }=C_f^{(u)}\mathrm{sin}\alpha \hfill \end{array}\}\text{for u-type fermions}.`$ (55)
Incorporating them into the decay amplitude yields:
$`A_{\mathrm{eff}}(hf\overline{f})`$ $`=`$ $`\sqrt{Z_h}\left(\mathrm{\Gamma }_h+Z_{hH}\mathrm{\Gamma }_H\right)`$ (56)
$`=`$ $`C_f^{(d)}\mathrm{cos}\mathrm{\Delta }\alpha \left(\mathrm{sin}\alpha \mathrm{tan}\mathrm{\Delta }\alpha \left(\mathrm{cos}\alpha \right)\right)`$
$`=`$ $`C_f^{(d)}\mathrm{sin}(\alpha +\mathrm{\Delta }\alpha )`$
$``$ $`C_f^{(d)}\mathrm{sin}\alpha _{\mathrm{eff}}\text{(for d-type fermions)}`$
$`A_{\mathrm{eff}}(hf\overline{f})`$ $``$ $`C_f^{(u)}\mathrm{cos}\alpha _{\mathrm{eff}}\text{(for u-type fermions)}`$ (57)
$`A_{\mathrm{eff}}(Hf\overline{f})`$ $``$ $`C_f^{(d)}\mathrm{cos}\alpha _{\mathrm{eff}}\text{(for d-type fermions)}`$ (58)
$`A_{\mathrm{eff}}(Hf\overline{f})`$ $``$ $`C_f^{(u)}\mathrm{sin}\alpha _{\mathrm{eff}}\text{(for u-type fermions)}`$ (59)
Recalling the relations
$$D(\alpha _{\mathrm{eff}})=D(\alpha )D(\mathrm{\Delta }\alpha )$$
(60)
and
$$\left(\begin{array}{cc}\widehat{\mathrm{\Sigma }}_H& \widehat{\mathrm{\Sigma }}_{hH}\\ \widehat{\mathrm{\Sigma }}_{hH}& \widehat{\mathrm{\Sigma }}_h\end{array}\right)=D^1(\alpha )\left(\begin{array}{cc}\widehat{\mathrm{\Sigma }}_{\varphi _1}& \widehat{\mathrm{\Sigma }}_{\varphi _1\varphi _2}\\ \widehat{\mathrm{\Sigma }}_{\varphi _1\varphi _2}& \widehat{\mathrm{\Sigma }}_{\varphi _2}\end{array}\right)D(\alpha )$$
(61)
it is obvious that $`\alpha _{\mathrm{eff}}=(\alpha +\mathrm{\Delta }\alpha `$) is exactly the angle that diagonalizes the higher order corrected Higgs boson mass matrix in the $`\varphi _1,\varphi _2`$-basis:
$`\left(\begin{array}{cc}m_{\varphi _1}^2\widehat{\mathrm{\Sigma }}_{\varphi _1}& m_{\varphi _1\varphi _2}^2\widehat{\mathrm{\Sigma }}_{\varphi _1\varphi _2}\\ m_{\varphi _1\varphi _2}^2\widehat{\mathrm{\Sigma }}_{\varphi _1\varphi _2}& m_{\varphi _2}^2\widehat{\mathrm{\Sigma }}_{\varphi _2}\end{array}\right)\stackrel{\alpha _{\mathrm{eff}}}{}\left(\begin{array}{cc}M_H^2& 0\\ 0& M_h^2\end{array}\right)`$ (66)
$`\alpha `$ (67)
$`\left(\begin{array}{cc}m_H^2\widehat{\mathrm{\Sigma }}_H& \widehat{\mathrm{\Sigma }}_{hH}\\ \widehat{\mathrm{\Sigma }}_{hH}& m_h^2\widehat{\mathrm{\Sigma }}_h\end{array}\right)\stackrel{\mathrm{\Delta }\alpha }{}\left(\begin{array}{cc}M_H^2& 0\\ 0& M_h^2\end{array}\right).`$ (72)
$`\alpha _{\mathrm{eff}}`$ can be obtained from
$$\alpha _{\mathrm{eff}}=\mathrm{arctan}\left[\frac{(M_A^2+M_Z^2)\mathrm{sin}\beta \mathrm{cos}\beta \widehat{\mathrm{\Sigma }}_{\varphi _1\varphi _2}}{M_Z^2\mathrm{cos}^2\beta +M_A^2\mathrm{sin}^2\beta \widehat{\mathrm{\Sigma }}_{\varphi _1}m_h^2}\right],\frac{\pi }{2}<\alpha _{\mathrm{eff}}<\frac{\pi }{2}.$$
(73)
### 2.2 The Higgs-boson propagator corrections
For the Higgs boson self-energies employed in eqs. (29)โ(35) we use the currently most accurate result based on Feynman-diagrammatic calculations. It contains the result of the complete one-loop on-shell calculation of Ref. , together with the dominant two-loop corrections of $`๐ช(\alpha \alpha _s)`$ obtained in Refs. , including also the leading terms of $`๐ช(G_F^2m_t^6/M_W^2)`$ ; the Fortran program FeynHiggs, based on this result, has been described in Ref. . In this way the complete MSSM one-loop on-shell result together with the dominant two-loop contribution, originating from the $`t\stackrel{~}{t}`$-sector (without any restrictions on the mixing), is taken into account in the propagator corrections.
In the approach in Refs. the Higgs boson self-energies are given by:
$$\widehat{\mathrm{\Sigma }}_s(q^2)=\widehat{\mathrm{\Sigma }}_s^{(1)}(q^2)+\widehat{\mathrm{\Sigma }}_s^{(2)}(0),s=h,H,hH,$$
(74)
where the momentum dependence has been neglected only at the two-loop level, while the full momentum dependence is kept in the one-loop contributions.
In a first step of approximation for the calculation of the decay width $`\mathrm{\Gamma }(hf\overline{f})`$ the momentum dependence is neglected everywhere in the Higgs boson self-energies (see eq. (36)):
$$\widehat{\mathrm{\Sigma }}_s(q^2)\widehat{\mathrm{\Sigma }}_s^{(1)}(0)+\widehat{\mathrm{\Sigma }}_s^{(2)}(0),s=h,H,hH.$$
(75)
This corresponds to the $`\alpha _{\mathrm{eff}}`$-approximation, as described in Sect. 2.1.
In a second step of approximation we approximate the Higgs boson self-energies by the compact analytical formulas given in Ref. :
$$\widehat{\mathrm{\Sigma }}_s(q^2)=\widehat{\mathrm{\Sigma }}_s^{(1)\mathrm{approx}}(0)+\widehat{\mathrm{\Sigma }}_s^{(2)\mathrm{approx}}(0),s=h,H,hH,$$
(76)
yielding relatively short expressions which allow a very fast numerical evaluation. In the following, this approximation is labeled by $`\alpha _{\mathrm{eff}}(\mathrm{approx})`$.
For the $`\stackrel{~}{t}`$-sector, we use the same conventions as in Ref. : the scalar top masses and the mixing angle are related to the parameters $`M_{\stackrel{~}{t}_L}`$, $`M_{\stackrel{~}{t}_R}`$ and $`X_t`$ of the $`\stackrel{~}{t}`$-mass matrix
$$_{\stackrel{~}{t}}^2=\left(\begin{array}{cc}M_{\stackrel{~}{t}_L}^2+m_t^2+\mathrm{cos}2\beta (\frac{1}{2}\frac{2}{3}s_W^2)M_Z^2& m_tX_t\\ m_tX_t& M_{\stackrel{~}{t}_R}^2+m_t^2+\frac{2}{3}\mathrm{cos}2\beta s_W^2M_Z^2\end{array}\right),$$
(77)
with
$$X_t=A_t\mu \mathrm{cot}\beta .$$
(78)
In the numerical analysis below we have chosen $`m_{\stackrel{~}{q}}M_{\stackrel{~}{t}_L}=M_{\stackrel{~}{t}_R}`$.
In Refs. it has been shown that for a given set of MSSM parameters the maximal values of $`M_h`$ as a function of $`X_t`$ are obtained for $`|X_t/m_{\stackrel{~}{q}}|2`$. This case we refer to as โmaximal mixingโ. Minimal values for $`M_h`$ are reached for $`X_t0`$. This case we refer to as โno mixingโ.
### 2.3 Decay width of the lightest Higgs boson
At the tree level, the decay width for $`hf\overline{f}`$ is given by
$$\mathrm{\Gamma }_0(hf\overline{f})=N_C\frac{m_h}{8\pi }\left(1\frac{4m_f^2}{m_h^2}\right)^{\frac{3}{2}}|\mathrm{\Gamma }_h|^2.$$
(79)
The electroweak propagator corrections are incorporated by using the higher-order decay amplitude (30)
$$\mathrm{\Gamma }_1\mathrm{\Gamma }_1(hf\overline{f})=N_C\frac{M_h}{8\pi }\left(1\frac{4m_f^2}{M_h^2}\right)^{\frac{3}{2}}|A(hf\overline{f})|^2.$$
(80)
The $`\alpha _{\mathrm{eff}}`$-approximation is given by
$$\mathrm{\Gamma }_{1,\mathrm{eff}}\mathrm{\Gamma }_{1,\mathrm{eff}}(hf\overline{f})=N_C\frac{M_h}{8\pi }\left(1\frac{4m_f^2}{M_h^2}\right)^{\frac{3}{2}}|A_{\mathrm{eff}}(hf\overline{f})|^2.$$
(81)
In this paper we consider only those electroweak higher-order contributions which enter via the Higgs boson self-energies. These corrections contain the Yukawa contributions of $`๐ช(G_Fm_t^4/M_W^2)`$, which are the dominant electroweak one-loop corrections to the Higgs-boson decay width, and the corresponding dominant two-loop corrections, see Sect. 2.2. The pure weak $`๐ช(\alpha )`$ vertex corrections are neglected (they have been calculated in Ref. and were found to be at the level of only a few % for most parts of the MSSM parameter space, see also Sect. 2.3.3 below.)
#### 2.3.1 QED corrections
Here we follow the results given in Refs. . The IR-divergent virtual photon contribution is taken into account in combination with real-photon bremsstrahlung yielding the QED corrections. The contribution to the decay width induced by $`\gamma `$-exchange and final-state photon radiation can be cast into the very compact formula
$$\mathrm{\Delta }\mathrm{\Gamma }_\gamma =\mathrm{\Gamma }_1\delta \mathrm{\Gamma }_\gamma ,$$
(82)
where for $`m_f^2M_h^2`$ the factor $`\delta \mathrm{\Gamma }_\gamma `$ has the simple form
$$\delta \mathrm{\Gamma }_\gamma =\frac{\alpha }{\pi }Q_f^2\left[3\mathrm{log}\left(\frac{M_h}{m_f}\right)+\frac{9}{4}\right].$$
(83)
#### 2.3.2 QCD corrections: gluon contributions
The corresponding results have been obtained in Refs. . The additional contribution to the decay width induced by gluon exchange and final-state gluon radiation can be incorporated into (80) by writing
$$\mathrm{\Gamma }_{1,g}=\mathrm{\Gamma }_1\frac{m_q^2(M_h^2)}{m_q^2}\left[1+\frac{\alpha _s(M_h^2)}{\pi }\left\{C_F\frac{9}{4}+\frac{8}{3}\left(1\frac{\alpha _s(m_b^2)}{\alpha _s(M_h^2)}\right)\right\}\right].$$
(84)
The correction factor containing $`\alpha _s(m_b^2)`$, which has not been included in previous diagrammatic calculations, can give rise to non-negligible contributions. $`m_q(M_h^2)`$ is calculated via
$`m_q(q^2)`$ $`=`$ $`m_q{\displaystyle \frac{c(q^2)}{c(m_q^2)}},`$ (85)
$`c(q^2)`$ $`=`$ $`\left({\displaystyle \frac{\beta _0\alpha _s(q^2)}{2\pi }}\right)^{\gamma _0/2\beta _0}[1+{\displaystyle \frac{\left(\beta _1\gamma _0\beta _0\gamma _1\right)}{\beta _0^2}}{\displaystyle \frac{\alpha _s(q^2)}{8\pi }}`$ (86)
$`+({\displaystyle \frac{(\beta _1\gamma _0\beta _0\gamma _1)^2}{2\beta _0^4}}+{\displaystyle \frac{\gamma _0(\beta _2\beta _0\beta _1^2)}{\beta _0^3}}+{\displaystyle \frac{\gamma _1\beta _1}{\beta _0^2}}{\displaystyle \frac{\gamma _2}{\beta _0}})\left({\displaystyle \frac{\alpha _s(q^2)}{8\pi }}\right)^2],`$
where $`m_q`$ is the pole mass and $`m_q(m_q^2)=m_q`$. The coefficients in eq. (86) are:
$`\beta _0`$ $`=`$ $`{\displaystyle \frac{332N_f}{3}},`$
$`\beta _1`$ $`=`$ $`102{\displaystyle \frac{38}{3}}N_f,`$
$`\beta _2`$ $`=`$ $`{\displaystyle \frac{2857}{2}}{\displaystyle \frac{5033}{18}}N_f+{\displaystyle \frac{325}{54}}N_f^2,`$
$`\gamma _0`$ $`=`$ $`8,`$
$`\gamma _1`$ $`=`$ $`{\displaystyle \frac{404}{3}}+{\displaystyle \frac{40}{9}}N_f,`$
$`\gamma _2`$ $`=`$ $`{\displaystyle \frac{2}{3}}\left[{\displaystyle \frac{140}{27}}N_f^2+\left(160\zeta (3)+{\displaystyle \frac{2216}{9}}\right)N_f3747\right],`$ (87)
where $`\zeta (3)1.2020596\mathrm{}`$ and $`N_f=5`$ for $`f=c,b`$, which are considered here. The strong coupling constant $`\alpha _s`$ is given up to three loops by:
$$\alpha _s(q^2)=\frac{4\pi }{\beta _0L_q}\left[1\frac{\beta _1}{\beta _0^2}\frac{\mathrm{log}L_q}{L_q}+\frac{\beta _1^2}{\beta _0^4}\frac{\mathrm{log}^2L_q}{L_q^2}\frac{\beta _1^2}{\beta _0^4}\frac{\mathrm{log}L_q}{L_q^2}+\frac{\beta _2\beta _0\beta _1^2}{\beta _0^4}\frac{1}{L_q^2}\right],$$
(88)
where $`L_q=\mathrm{log}(q^2/\mathrm{\Lambda }_{\mathrm{QCD}}^2)`$. (For the numerical evaluation $`\mathrm{\Lambda }_{\mathrm{QCD}}=220\mathrm{MeV}`$ has been used.) Numerically, more than 80% of the gluon-exchange contribution is absorbed into the running quark mass.
#### 2.3.3 QCD corrections: gluino contributions
We follow the calculation given in Ref. <sup>1</sup><sup>1</sup>1 An error in Ref. concerning the proper inclusion of the $`Hf\overline{f}`$ coupling has been corrected. , similar results can also be found in Ref. . The additional contributions to the decay width induced by gluino-exchange are incorporated via
$$\mathrm{\Delta }\mathrm{\Gamma }_{\stackrel{~}{g}}=\mathrm{\Gamma }_1\delta \mathrm{\Gamma }_{\stackrel{~}{g}},$$
(89)
where $`\delta \mathrm{\Gamma }_{\stackrel{~}{g}}`$ is given by
$$\delta \mathrm{\Gamma }_{\stackrel{~}{g}}=\frac{2}{\mathrm{\Gamma }_h+Z_{hH}\mathrm{\Gamma }_H}\text{Re}\left[\mathrm{\Gamma }_{\stackrel{~}{g}}^h+Z_{hH}\mathrm{\Gamma }_{\stackrel{~}{g}}^H\right]$$
(90)
for real $`Z_{hH}`$ (i.e. neglecting the imaginary part in eq. (90)); $`\mathrm{\Gamma }_{\stackrel{~}{g}}^h`$ and $`\mathrm{\Gamma }_{\stackrel{~}{g}}^H`$ are given by
$`\mathrm{\Gamma }_{\stackrel{~}{g}}^h`$ $`=`$ $`\mathrm{\Gamma }_h[\mathrm{\Delta }T_{\stackrel{~}{g}}^h{}_{|q^2=M_h^2}{}^{}+\mathrm{\Sigma }_{S,\stackrel{~}{g}}^f(m_q^2)2m_q^2(\mathrm{\Sigma }_{S,\stackrel{~}{g}}^f(m_q^2)+\mathrm{\Sigma }_{V,\stackrel{~}{g}}^f(m_q^2))]`$ (91)
$`\mathrm{\Gamma }_{\stackrel{~}{g}}^H`$ $`=`$ $`\mathrm{\Gamma }_H[\mathrm{\Delta }T_{\stackrel{~}{g}}^H{}_{|q^2=M_h^2}{}^{}+\mathrm{\Sigma }_{S,\stackrel{~}{g}}^f(m_q^2)2m_q^2(\mathrm{\Sigma }_{S,\stackrel{~}{g}}^f(m_q^2)+\mathrm{\Sigma }_{V,\stackrel{~}{g}}^f(m_q^2))].`$ (92)
$`\mathrm{\Delta }T_{\stackrel{~}{g}}^{h,H}`$ denote the gluino vertex-corrections, whereas $`\mathrm{\Sigma }^f`$ represents the gluino contribution to the fermion self-energy corrections. Explicit expressions for these terms can be found in Ref. .
For large values of $`\mathrm{tan}\beta `$ in combination with large values of $`|\mu |`$, the gluino-exchange corrections to $`\mathrm{\Gamma }(hb\overline{b})`$ can become very large. In Refs. as well as in Ref. it has been proposed to derive an effective contribution to the decay width resummed to all orders. A similar resummation can be applied for weak $`๐ช(\alpha )`$ chargino-exchange corrections to $`\mathrm{\Gamma }(hb\overline{b})`$, where they become non-negligible . In the numerical examples given in Ref. the difference between the one-loop result for $`\mathrm{\Gamma }(hf\overline{f})`$ and the effectively resummed result does not exceed 10 โ 15%, even for very large values of $`\mathrm{tan}\beta >40`$. A proof of how the resummation of the leading terms arising for large $`\mathrm{tan}\beta `$ can be performed for the $`H^+\overline{t}b`$ vertex is given in Ref. . Concerning our numerical analysis in Sect. 3 we have neglected the additional contributions from the resummation. These additional corrections, although potentally large in some regions of the MSSM parameter space, would not qualitatively change our conclusions given below. A more detailed investigation of the effects of a proper resummation of the leading contributions and of the inclusion of the complete electroweak one-loop vertex corrections will be given in a forthcoming publication.
For some parameter combinations the gluino corrections can drive $`\mathrm{\Gamma }(hb\overline{b})`$ to very small values, see the discussion at the end of Sect. 3.1.
#### 2.3.4 Decay width and branching ratio
Including the various types of corrections, the decay width is given by
$$\mathrm{\Gamma }(hf\overline{f})=\mathrm{\Gamma }_{1,g}+\mathrm{\Delta }\mathrm{\Gamma }_\gamma +\mathrm{\Delta }\mathrm{\Gamma }_{\stackrel{~}{g}}.$$
(93)
Summing over $`f=b,c,\tau `$ and adding $`\mathrm{\Gamma }(hgg)`$ (which can be numerically relevant ), results in an approximation for the total decay width
$$\mathrm{\Gamma }_{\mathrm{tot}}=\underset{f=b,c,\tau }{}\mathrm{\Gamma }(hf\overline{f})+\mathrm{\Gamma }(hgg).$$
(94)
We do not take into account the decay $`hAA`$ (see e.g. Ref. for a detailed study). Although it is dominant whenever it is kinematically allowed, it plays a role only for very small values of $`\mathrm{tan}\beta `$ ($`\mathrm{tan}\beta \stackrel{<}{}\mathrm{\hspace{0.33em}1.5}`$) which will not be considered here because of the limits obtained at LEP2 . We also assume that all other SUSY particles are too heavy to allow further decay channels. In addition, we neglect the decay $`hWW^{}`$ which can be of $`๐ช(1\%)`$ for $`M_h\stackrel{>}{}\mathrm{\hspace{0.33em}100}\mathrm{GeV}`$.
The fermionic branching ratio is defined by
$$R_fBR(hf\overline{f})=\frac{\mathrm{\Gamma }(hf\overline{f})}{\mathrm{\Gamma }_{\mathrm{tot}}}.$$
(95)
## 3 Numerical analysis
Concerning the numerical evaluation of the Higgs-boson propagator corrections, we follow Sect. 2.2. For $`\mathrm{tan}\beta `$ we have chosen two representative values, a relatively low value, $`\mathrm{tan}\beta =3`$,<sup>2</sup><sup>2</sup>2 This values is well above the expected limit obtainable at the end of LEP, assuming that no Higgs boson signal will be found . For these expected limits $`m_t=174.3\mathrm{GeV}`$ and $`m_{\stackrel{~}{q}}=1\mathrm{TeV}`$ has been assumed. and a high value, $`\mathrm{tan}\beta =40`$. For sake of comparison we also consider an intermediate value of $`\mathrm{tan}\beta =20`$ in some cases. If not indicated differently, the other MSSM parameters are chosen as follows: $`\mu =100\mathrm{GeV}`$, $`M_2=m_{\stackrel{~}{q}}`$ ($`M_2`$ is the soft SUSY-breaking term in the gaugino sector), gluino mass $`m_{\stackrel{~}{g}}=500\mathrm{GeV}`$, $`A_b=A_t`$ (which fixes, together with $`\mu `$, the mixing in the $`\stackrel{~}{b}`$-sector). For the SM fermion masses we have furthermore chosen $`m_t=175\mathrm{GeV}`$, $`m_b=4.5\mathrm{GeV}`$<sup>3</sup><sup>3</sup>3 The value of $`m_h=4.5\mathrm{GeV}`$ is used in the tree-level expression and in the QED and QCD vertex corrections (see Sects. 2.3.1 \- 2.3.3), while for the Higgs-propagator corrections the running bottom mass, $`m_b(m_t)=2.97\mathrm{GeV}`$, has been used, in order to partially absorb higher-order QCD corrections. , $`m_\tau =1.777\mathrm{GeV}`$ and $`m_c=1.5\mathrm{GeV}`$.
The mass $`M_A`$ of the $`๐๐ซ`$-odd Higgs boson is treated as an input parameter and is varied in the interval $`50\mathrm{GeV}M_A500\mathrm{GeV}`$. The corresponding values for $`M_h`$ follow from eq. (29). $`M_h`$, derived in this way, subsequently enters the numerical evaluation of the formulas presented in section 2. Thus the variation of $`M_h`$ in the plots stems from the variation of $`M_A`$ in the above given range.
### 3.1 Effects of the two-loop Higgs-propagator corrections
We first focus on the effects of the two-loop Higgs-boson propagator corrections. They have been evaluated at the one- and at the two-loop level as described in Sect. 2.2. Fig. 1 shows the results for $`\mathrm{\Gamma }(hb\overline{b})`$ for a common scalar quark mass $`m_{\stackrel{~}{q}}=1000\mathrm{GeV}`$ and $`\mathrm{tan}\beta =3`$ and $`\mathrm{tan}\beta =40`$ in the no-mixing and the maximal-mixing scenario. The QED and the QCD gluon and gluino vertex contributions are also included.
In the small $`\mathrm{tan}\beta `$ scenario, larger values for $`\mathrm{\Gamma }(hb\overline{b})`$ are obtained for maximal mixing. The two-loop corrections strongly reduce the decay width. In the large $`\mathrm{tan}\beta `$ scenario the variation is mainly a kinematical effect from the different values of $`M_h`$ at the one- and two-loop level. The absolute values obtained for $`\mathrm{\Gamma }(hb\overline{b})`$ are three orders of magnitude higher in the $`\mathrm{tan}\beta =40`$ scenario, which is due to the fact that $`\mathrm{\Gamma }(hb\overline{b})1/\mathrm{cos}^2\beta `$.
In Fig. 2 the three decay rates $`\mathrm{\Gamma }(hb\overline{b})`$, $`\mathrm{\Gamma }(h\tau ^+\tau ^{})`$ and $`\mathrm{\Gamma }(hc\overline{c})`$ are shown as a function of $`M_h`$. The results are given in the no-mixing scenario for $`m_{\stackrel{~}{q}}=500\mathrm{GeV}`$ and $`\mathrm{tan}\beta =3,40`$.
In the low $`\mathrm{tan}\beta `$ scenario $`\mathrm{\Gamma }(hb\overline{b})`$ and $`\mathrm{\Gamma }(h\tau ^+\tau ^{})`$ are lowered at the two-loop level, while $`\mathrm{\Gamma }(hc\overline{c})`$ is increased. The decay rate for $`hb\overline{b}`$ is about one and two orders of magnitude larger compared to the ones of $`h\tau ^+\tau ^{}`$ and $`hc\overline{c}`$, respectively. In the large $`\mathrm{tan}\beta `$ scenario the shifts are again dominated by the kinematical effect from the different values of $`M_h`$ at the one- and two-loop level. In the maximal-mixing case, which is not plotted here, we find qualitatively the same behavior.
We now turn to the effects of the two-loop corrections to the branching ratios. For not too large values of $`M_h`$, $`\mathrm{\Gamma }_{\mathrm{tot}}`$ is strongly dominated by $`\mathrm{\Gamma }(hb\overline{b})`$. For large values of $`M_h`$ the decay into gluons becomes more relevant. In Fig. 3 we show the branching ratio $`BR(hb\overline{b})`$ as a function of $`M_h`$ and $`M_A`$. For values of $`M_A\stackrel{>}{}\mathrm{\hspace{0.33em}250}\mathrm{GeV}`$ there is a non-negligible difference between one-loop and two-loop order, where at the two-loop level the branching ratio is slightly enhanced. Compared in terms of $`M_h`$ there is nearly no change for small values of $`M_h`$ in the low and in the high $`\mathrm{tan}\beta `$ case. Here $`BR(hb\overline{b})`$ is changed by less than about 1%, see Fig. 3. $`BR(h\tau ^+\tau ^{})`$ is increased by less than about 2%. $`BR(hc\overline{c})`$ can be increased at the two-loop level by $`๐ช(50\%)`$, but remains numerically relatively small. For $`\mathrm{tan}\beta =40`$ the main difference arises at the endpoints of the spectrum, again due to the fact that different Higgs boson masses can be obtained at the one-loop and at the two-loop level. For $`\mathrm{tan}\beta =3`$, however, also several GeV below the kinematical endpoints there is a sizable effect on $`BR(hb\overline{b})`$. Thus, in the experimentally allowed region of $`M_h`$, the two-loop corrections can have an important effect on $`BR(hb\overline{b})`$.
Higgs boson search, especially at $`e^+e^{}`$ colliders, often relies on $`b`$ search, since on one hand the lightest $`๐๐ซ`$-even Higgs boson decays dominantly into $`b\overline{b}`$ and on the other hand $`b`$ tagging can be performed with high efficiency. For some combinations of parameters, however, $`\mathrm{\Gamma }(hb\overline{b})`$ can become very small and thus $`BR(hb\overline{b})`$ can approach zero as a consequence of large Higgs-boson propagator corrections or large gluino vertex-corrections, making Higgs boson search possibly very difficult for these parameters. Within the effective potential approach this kind of effect has first been observed in Ref. , recent analyses investigating the parameter regions where $`BR(hb\overline{b})`$ is suppressed can be found in Refs. . In order to have reliable predictions for these regions of parameter space a full calculation of the one-loop vertex corrections, including all $`๐ช(\alpha )`$ contributions, would be necessary. Here we demonstrate the effect of the two-loop propagator corrections on the values of the parameters, especially of $`M_A`$, for which $`BR(hb\overline{b})`$ goes to zero. We also show the impact of the inclusion of the momentum dependence of the Higgs boson self-energies (see eq. (29) and (36)), that is often neglected in phenomenological analyses of the decays of the lightest $`๐๐ซ`$-even Higgs boson.
In Fig. 4 $`BR(hb\overline{b})`$ is shown as a function of $`M_A`$. The Higgs boson self-energies are evaluated at the one-loop and at the two-loop level with and without momentum dependence (see eq. (36)). The other parameters are $`\mathrm{tan}\beta =25`$, $`m_{\stackrel{~}{q}}=500\mathrm{GeV}`$, $`m_{\stackrel{~}{g}}=400\mathrm{GeV}`$, $`M_2=400\mathrm{GeV}`$, $`X_t=400\mathrm{GeV}`$, $`A_b=A_t`$, $`\mu =1000\mathrm{GeV}`$. The inclusion of the two-loop propagator corrections shifts the $`M_A`$ value for which $`BR(hb\overline{b})`$ becomes very small by about $`35\mathrm{GeV}`$. The inclusion of the momentum dependence of the Higgs boson self-energies induces another shift of about $`6\mathrm{GeV}`$. In order to have reliable phenomenological predictions for the problematic $`M_A`$ values the two-loop corrections as well as the inclusion of the momentum dependence is necessary. Note that the inclusion of the gluino vertex corrections as well as the purely weak vertex corrections can also have a large impact on the critical $`M_A`$ values.
### 3.2 Effects of the gluino vertex corrections
In this subsection we present the effect of the gluino-exchange contribution to the $`hf\overline{f}`$ vertex corrections. These corrections have been neglected so far in most phenomenological analyses<sup>4</sup><sup>4</sup>4 The gluino-exchange contributions are currently incorporated into HDECAY . Concerning the discovery potential of LEP, the Tevatron and the LHC, the gluino corrections have also been studied recently in Refs. . .
Fig. 5 shows $`\mathrm{\Gamma }(hb\overline{b})`$ in three steps of accuracy: the dotted curves contain only the pure self-energy correction, the dashed curves contain in addition the QED and the gluon-exchange correction. The solid curves show the full results, including also the gluino-exchange correction. The results are shown for the no mixing scenario, $`\mu =100\mathrm{GeV}`$, $`m_{\stackrel{~}{g}}=500\mathrm{GeV}`$, $`m_{\stackrel{~}{q}}=200,1000\mathrm{GeV}`$ in the left and right part of Fig. 5, respectively. For $`\mathrm{tan}\beta `$, three values have been chosen: $`\mathrm{tan}\beta =3,20,40`$.
The left plot of Fig. 5 corresponds to a small soft SUSY-breaking scale, $`m_{\stackrel{~}{q}}=200\mathrm{GeV}`$. The effect of the gluon contribution is large and negative, the effect of the QED correction is small. For this combination of $`m_{\stackrel{~}{g}}`$, $`m_{\stackrel{~}{q}}`$ and $`\mu `$ the effect of the gluino correction is large and positive as can be seen from the transition from the dashed to the solid curves. For $`\mathrm{tan}\beta =40`$ it nearly compensates the gluon effect, for $`\mathrm{tan}\beta =20`$ it amounts up to 20% of the gluonic correction, while for $`\mathrm{tan}\beta =3`$ the gluino-exchange contribution is negligible. Note that we have chosen a relatively small value of $`\mu `$, $`\mu =100\mathrm{GeV}`$. For larger values of $`|\mu |`$ even larger correction can be obtained. Hence neglecting the gluino-exchange correction in the large $`\mathrm{tan}\beta `$ scenario can lead to results which deviate by 50% from the full $`๐ช(\alpha _s)`$ calculation (see also Sect. 2.3.3). The right plot of Fig. 5 corresponds to $`m_{\stackrel{~}{q}}=1000\mathrm{GeV}`$. The gluino-exchange effects are still visible, but much smaller than for $`m_{\stackrel{~}{q}}=200\mathrm{GeV}`$. The same observation has already been made in Ref. . In the maximal-mixing scenario we find qualitatively the same behavior for the gluino-exchange corrections as in the no-mixing scenario.
In Fig. 6 the pure gluino-exchange effect is shown as a function of $`\mu `$. This effect increases with rising<sup>5</sup><sup>5</sup>5 This is correct for all values of $`m_{\stackrel{~}{g}}`$ considered in this work. A maximal effect is reached around $`m_{\stackrel{~}{g}}1500\mathrm{GeV}`$. The decoupling of the gluino takes place only for very large values, $`m_{\stackrel{~}{g}}\stackrel{>}{}\mathrm{\hspace{0.33em}5000}\mathrm{GeV}`$. $`m_{\stackrel{~}{g}}`$ and $`|\mu |`$, where for negative (positive) $`\mu `$ there is an enhancement (a decrease) in $`\mathrm{\Gamma }(hb\overline{b})`$. The size of the gluino-exchange contribution also depends on $`M_A`$, where larger effects correspond to smaller values of $`M_A`$, see also Ref. . The small difference between the curves where the decay rate has been calculated without gluino contribution is due to the variation of $`M_h`$ induced by different values of $`m_{\stackrel{~}{g}}`$ which enters at $`๐ช(\alpha \alpha _s)`$. Fig. 6 demonstrates again that neglecting the gluino contribution in the fermion decay rates can yield (strongly) misleading results.
The gluino exchange contribution has only a relatively small impact on $`BR(hb\overline{b})`$. It can have a large influence, on the other hand, on $`BR(h\tau ^+\tau ^{})`$. Both branching ratios are expected to be measurable at the same level of accuracy, see e.g. Ref. . While the Higgs-propagator contributions are universal corrections that affect $`\mathrm{\Gamma }(hb\overline{b})`$ and $`\mathrm{\Gamma }(h\tau ^+\tau ^{})`$ in the same way (i.e. the influence on the effective coupling is the same in both cases), the gluino corrections, which influence only $`\mathrm{\Gamma }(hb\overline{b})`$, can lead to a different behavior of the two decay widths. In Fig. 7 we show $`BR(h\tau ^+\tau ^{})`$ as a function of $`m_{\stackrel{~}{g}}`$. The left plot corresponds to three different values of $`\mathrm{tan}\beta `$ and $`\mu =\pm 100\mathrm{GeV}`$, $`M_A=100\mathrm{GeV}`$. Here we have furthermore chosen $`m_{\stackrel{~}{q}}=500\mathrm{GeV}`$ and moderate mixing, i.e. $`X_t=m_{\stackrel{~}{q}}`$; we find similar results for the other mixing scenarios. For positive (negative) $`\mu `$ the decay rate $`\mathrm{\Gamma }(hb\overline{b})`$ is reduced (enhanced), see Fig. 6, thus $`BR(h\tau ^+\tau ^{})`$ is increased (decreased) by up to 50%. For large values of $`m_{\stackrel{~}{g}}`$ and $`\mathrm{tan}\beta `$, $`BR(h\tau ^+\tau ^{})`$ can thus be considerably different from the case where the gluino-exchange contribution to $`\mathrm{\Gamma }(hb\overline{b})`$ has been neglected. This becomes even more apparent in the right plot of Fig. 7, where the MSSM result, including gluino exchange contribution, is compared to the SM result. Here we also show the scenario with $`M_A=300\mathrm{GeV}`$ and $`\mu =400\mathrm{GeV}`$, where the Higgs sector of the MSSM behaves SM like (i.e. the lightest Higgs boson has almost SM couplings, all other Higgs bosons are heavy). The horizontal lines represent the SM values for the respective Higgs boson masses. The two Higgs boson masses for each line give similar results so that the lines are indistinguishable in the plot. These masses correspond to an averaged value obtained in the MSSM in the interval $`0<m_{\stackrel{~}{g}}<1000\mathrm{GeV}`$, where the variation of $`M_h`$ is about $`\pm 1\mathrm{GeV}`$ (with our choice of $`m_{\stackrel{~}{q}}`$ and $`X_t`$). Since the gluino decouples very slowly, there is no decoupling effect for $`m_{\stackrel{~}{g}}\stackrel{<}{}\mathrm{\hspace{0.33em}1000}\mathrm{GeV}`$ and the MSSM results can be considerably different from the SM result, for $`M_A=100\mathrm{GeV}`$ as well as for $`M_A=300\mathrm{GeV}`$, i.e. even where the MSSM Higgs sector behaves otherwise SM like. The deviation can amount up to 50% for $`M_A=`$ $`๐ช(100\mathrm{GeV})`$ and up to 30% for $`M_A=`$ $`๐ช(300\mathrm{GeV})`$. Thus the measurement of $`BR(h\tau ^+\tau ^{})`$ can provide a distinction between the SM and the MSSM even for relatively large $`M_A`$ and large $`m_{\stackrel{~}{g}}`$ if also $`|\mu |`$ and $`\mathrm{tan}\beta `$ are sufficiently large. The branching ratio $`BR(hb\overline{b})`$ on the other hand, changes only by a few per cent for these parameters, so that this change would be much harder to measure.
### 3.3 The $`\alpha _{\mathrm{eff}}`$-approximation
In this section we investigate the quality of the $`\alpha _{\mathrm{eff}}`$-approximation. In Fig. 8 we display the relative difference between the full result (74) and the $`\alpha _{\mathrm{eff}}`$ result, where the external momentum of the Higgs self-energies has been neglected, see eq. (75). The relative difference $`\mathrm{\Delta }\mathrm{\Gamma }(hb\overline{b})=(\mathrm{\Gamma }^{\mathrm{full}}(hb\overline{b})\mathrm{\Gamma }^{\alpha _{\mathrm{eff}}}(hb\overline{b}))/\mathrm{\Gamma }^{\mathrm{full}}(hb\overline{b})`$ is shown as a function of $`M_A`$ for $`m_{\stackrel{~}{q}}=1000\mathrm{GeV}`$ and for three values of $`\mathrm{tan}\beta `$ in the no mixing and the maximal mixing scenario. Large deviations occur only in the region $`100\mathrm{GeV}\stackrel{<}{}M_A\stackrel{<}{}\mathrm{\hspace{0.33em}150}\mathrm{GeV}`$, especially for large $`\mathrm{tan}\beta `$. In this region of parameter space the values of $`M_h`$ and $`M_H`$ are very close to each other. This results in a high sensitivity to small deviations in the Higgs boson self-energies entering the Higgs-boson mass matrix (29), (37). Another source of differences between the full and the approximate calculation is the threshold $`M_A=2m_t=350\mathrm{GeV}`$ in the one-loop contribution, originating from the top-loop diagram in the $`A`$ self-energy and in the $`AZ`$ self-energy (see Ref. ). Here the deviation can amount up to 6%.
In Fig. 9 we compare the $`\alpha _{\mathrm{eff}}`$ result (75) with the $`\alpha _{\mathrm{eff}}(\mathrm{approx})`$ result (76), where the Higgs boson self-energies have been approximated by the compact analytical expression obtained in Ref. . Fig. 9 displays the relative difference in the effective mixing angles, $`(\mathrm{sin}\alpha _{\mathrm{eff}}\mathrm{sin}\alpha _{\mathrm{eff}}(\mathrm{approx}))/\mathrm{sin}\alpha _{\mathrm{eff}}`$. Via eq. (56) $`\mathrm{sin}\alpha _{\mathrm{eff}}`$ directly determines the decay width $`\mathrm{\Gamma }(hb\overline{b})`$. The result is shown for $`m_{\stackrel{~}{q}}=1000\mathrm{GeV}`$, for three values of $`\mathrm{tan}\beta `$ in the minimal and the maximal mixing scenario. Apart from the region around $`M_A120\mathrm{GeV}`$ (compare Fig. 8) both effective angles agree better than 3% with each other.
Concerning the comparison of the $`\alpha _{\mathrm{eff}}`$-approximations in terms of $`M_h`$ (which is not plotted here), due to the neglected external momentum or the neglected subdominant one- or two-loop terms, $`M_h`$ receives a slight shift. Besides this kinematical effect, the decay rate is approximated rather well for most of the $`M_h`$ values: independently of $`m_{\stackrel{~}{q}}`$, the differences stay mostly below 2-4%, for the no-mixing case as well as for the maximal-mixing case. Only at the endpoints of the spectrum, due to the different Higgs-boson mass values, the difference is not negligible.
## 4 Comparison with the renormalization group approach
In order to compare our results with those obtained within the renormalization-group-improved effective field theory approach (in the following, for brevity reasons, denoted as RG approach), we made use of the program HDECAY . In Fig. 10 we show the decay rate $`\mathrm{\Gamma }(hb\overline{b})`$ as a function of $`M_h`$ (left part) and as a function of $`M_A`$ (right part). We compare the RG result of HDECAY, where the gluino-exchange contribution is not yet implemented, with the FD result without and with gluino correction. In order to trace back the source of deviations we also show in the lower part of Fig. 10 the sine of the effective mixing angle $`\alpha _{\mathrm{eff}}`$ which enters the decay rate $`\mathrm{\Gamma }(hb\overline{b})`$ quadratically (see eqs. (56) and (81)). In order to evaluate $`\alpha _{\mathrm{eff}}`$ we have neglected the external momentum (see eq. (75)). As a typical soft SUSY-breaking scale we have chosen $`m_{\stackrel{~}{q}}=500\mathrm{GeV}`$ in Fig. 10. Since the FD and the RG result have been obtained in different renormalization schemes, the entries of the $`\stackrel{~}{t}`$-mass matrix have a different meaning starting from two-loop order . For the no-mixing case we have set $`X_t=0`$ in both approaches, whereas the maximal-mixing case is defined via $`X_t=2m_{\stackrel{~}{q}}`$ for the FD approach and $`X_t=\sqrt{6}m_{\stackrel{~}{q}}`$ in the RG approach .
In the comparison of the decay rates (and restricting ourselves to the results with neglected gluino exchange), for a given $`M_h`$ we find deviations for low $`\mathrm{tan}\beta `$ up to $`๐ช(10\%)`$ and agreement better than 4% in the large $`\mathrm{tan}\beta `$ scenario. The main part of the deviations can be attributed to the deviations in $`\alpha _{\mathrm{eff}}`$ which modify $`\mathrm{\Gamma }(hb\overline{b})`$ (see the lower part of Fig. 10). As a general feature, larger deviations arise at the endpoints of the $`M_h`$ spectrum due to the fact that for the same value of $`M_A`$ different values for $`M_h`$ are obtained in the FD and the RG approach, as shown in the left part of Fig. 10. The plots for the $`\mathrm{tan}\beta =40`$ scenario show again a sizable effect of the gluino correction.
In Figs. 11 and Fig. 12 the relative differences $`\mathrm{\Delta }\mathrm{\Gamma }(hb\overline{b})=(\mathrm{\Gamma }^{\mathrm{FD}}(hb\overline{b})\mathrm{\Gamma }^{\mathrm{RG}}(hb\overline{b}))/\mathrm{\Gamma }^{\mathrm{FD}}(hb\overline{b})`$ and $`(\mathrm{sin}\alpha _{\mathrm{eff}}^{\mathrm{FD}}\mathrm{sin}\alpha _{\mathrm{eff}}^{\mathrm{RG}})/\mathrm{sin}\alpha _{\mathrm{eff}}^{\mathrm{FD}}`$ are shown as a function of $`M_A`$. Comparing $`\alpha _{\mathrm{eff}}`$ as a function of $`M_A`$ the agreement is relatively good; sizable deviations larger than 10% are found only in the regions where the curves in the right part of Fig. 10 have a steep slope. This can give rise to large deviations up to 50% in $`\mathrm{\Gamma }(hb\overline{b})`$ in terms of $`M_A`$ (for $`100\mathrm{GeV}\stackrel{<}{}M_A\stackrel{<}{}\mathrm{\hspace{0.33em}150}\mathrm{GeV}`$), as displayed in the right parts of Figs. 11 and 12 (large $`\mathrm{tan}\beta `$).
## 5 Conclusions
Using the Feynman-diagrammatic approach for the Higgs-boson propagator corrections, including besides the full one-loop result also the dominant two-loop corrections, we have calculated the decay rates and branching ratios for the decays $`hb\overline{b}`$, $`hc\overline{c}`$ and $`h\tau ^+\tau ^{}`$. We have included the one-loop QED and QCD corrections, where the latter are due to gluon and gluino-exchange contributions. The gluino-exchange corrections have been neglected in most of the previous phenomenological analyses. In the present analysis, only the purely weak $`๐ช(\alpha )`$ (process specific) vertex corrections, shown to contribute less than 1% for most parts of the MSSM parameter space, have been neglected.
We have shown analytically that the full set of Higgs-boson propagator corrections for vanishing external momentum can be absorbed into the effective mixing angle, $`\alpha _{\mathrm{eff}}`$, in the neutral $`๐๐ซ`$-even Higgs sector, appearing in the Higgs-boson fermion couplings.
Numerically we have shown that, compared in terms of $`M_h`$, the two-loop contributions to the Higgs-boson propagator corrections lead to a sizable decrease for $`\mathrm{\Gamma }(hb\overline{b})`$ and $`\mathrm{\Gamma }(h\tau ^+\tau ^{})`$, whereas $`\mathrm{\Gamma }(hc\overline{c})`$ can be increased, although it stays relatively small for all sets of parameters we have investigated. A sizable difference in all analyses from the one- to the two-loop calculation arises from the kinematical effect that at the two-loop level lower values for $`M_h`$ are obtained compared to the one-loop case, thus leading to deviations at the endpoints of the shown $`M_h`$ spectra. For most parts of the MSSM parameter space the $`\alpha _{\mathrm{eff}}`$-approximation reproduces the full result better than 3%. The gluino-exchange contribution to $`\mathrm{\Gamma }(hb\overline{b})`$ has been shown to be sizable for large $`\mathrm{tan}\beta `$. This correction increases with rising $`m_{\stackrel{~}{g}}`$ and decreases with rising $`m_{\stackrel{~}{q}}`$ and $`M_A`$. It is positive (negative) for negative (positive) $`\mu `$. In the $`\mathrm{tan}\beta =40`$ scenario for small $`m_{\stackrel{~}{q}}`$ it nearly compensate the gluon-exchange correction.
The effect of the Higgs-propagator corrections on the branching ratios $`BR(hf\overline{f})`$ is relatively small for small values of $`M_h`$, while sizable effects are possible in the experimentally favored region of $`M_h`$. The branching ratio $`BR(h\tau ^+\tau ^{})`$ can receive large corrections up to 50% due to SUSY-QCD corrections to $`\mathrm{\Gamma }(hb\overline{b})`$. In some parameter regions the effective $`hb\overline{b}`$ coupling can become very small, and $`BR(hb\overline{b})`$ can approach zero. In these parameter regions the two-loop contributions as well as the effect of the momentum dependence of the one-loop contributions are particularly important. It should be noted that in order to determine $`BR(hb\overline{b})`$ very precisely in these paremeter regions, also a resummation of the leading terms as well as the inclusion of the complete $`๐ช(\alpha )`$ vertex corrections will be necessary.
We have compared our diagrammatic result for $`\mathrm{\Gamma }(hb\overline{b})`$ with the result obtained within the renormalization-group-improved effective field theory approach. For the low $`\mathrm{tan}\beta `$ scenario, compared in terms of $`M_h`$, we find deviations up to $`๐ช(10\%)`$. In the large $`\mathrm{tan}\beta `$ scenario the agreement is better than 4%. Compared in terms of $`M_A`$, however, differences of up to 50% are possible in the region $`100\mathrm{GeV}\stackrel{<}{}M_A\stackrel{<}{}\mathrm{\hspace{0.33em}150}\mathrm{GeV}`$. The main part of the deviations can be attributed to the differences in the effective mixing angle $`\alpha _{\mathrm{eff}}`$.
### Acknowledgments
We thank M. Spira for valuable discussions and communication about the numerical comparison of our results. We are also grateful to M. Carena, H. Eberl and C. Wagner for interesting and helpful discussions. Parts of the calculations have been performed on the QCM cluster at the University of Karlsruhe. |
warning/0003/hep-ph0003325.html | ar5iv | text | # References
## Figure captions
Figure 1. The distribution functions $`F_B^b(x)`$ (solid line) and $`F_{\mathrm{\Lambda }_b}^b(x)`$ (thin lines) defined by Eq. (7) versus the LF momentum fraction $`x`$. Labels 1 to 4 on the curves refer to the cases $`\beta _{bu}=0.3\mathrm{GeV},m_{ud}=600\mathrm{MeV}`$; $`\beta _{bu}=0.3\mathrm{GeV},m_{ud}=300\mathrm{MeV}`$; $`\beta _{bu}=0.6\mathrm{GeV},m_{ud}=600\mathrm{MeV}`$; $`\beta _{bu}=0.6\mathrm{GeV},m_{ud}=300\mathrm{MeV}`$, respectively.
Figure 2. The lifetime ratios $`\tau _{\mathrm{\Lambda }_b}/\tau _B`$ for $`\beta =0.3\mathrm{GeV}(r=0.3))`$ (solid line), $`\beta =0.45\mathrm{GeV}(r=1)`$ (longโdashed line), and $`\beta =0.6\mathrm{GeV}(r=2.37)`$ (shortโdashed line). (a) $`m_{ud}=m_u+m_d`$, and (b) $`m_{ud}\frac{1}{2}(m_u+m_dm_\pi )`$.
Figure 1
Figure 2 |
warning/0003/hep-ph0003181.html | ar5iv | text | # Precision Electroweak Parameters and the Higgs MassTo be published in the Proceedings of MuMu99-5th International Conference on Physics Potential and Development of ๐^ยฑโข๐โป Colliders, San Francisco, CA, Dec. (1999)
## Fundamental Parameters and Precision Measurements
The SU(2)$`{}_{L}{}^{}\times `$U(1)<sub>Y</sub> electroweak sector of the standard model contains 17 or more fundamental parameters. They include gauge and Higgs field couplings as well as fermion masses and mixing angles. In terms of those parameters, predictions can be made with high accuracy for essentially any electroweak observable. Very precise measurements of those quantities can then be used to test the standard model, at the quantum loop level and predict the Higgs scalar mass or search for small deviations from expectations which would indicate โNew Physicsโ.
Some fundamental electroweak parameters have been determined with extraordinary precision. Foremost in that category is the fine structure constant $`\alpha `$. It is best obtained by comparing the measured prl5926 anomalous magnetic moment of the electron, $`a_e(g_e2)/2`$
$$a_e^{\mathrm{exp}}=1159652188(3)\times 10^{12}$$
(1)
with the calculated 4 loop QED prediction hepph9810512
$`a_e^{\mathrm{th}}`$ $`=`$ $`{\displaystyle \frac{\alpha }{2\pi }}0.328478444\left({\displaystyle \frac{\alpha }{\pi }}\right)^2+1.181234\left({\displaystyle \frac{\alpha }{\pi }}\right)^31.5098\left({\displaystyle \frac{\alpha }{\pi }}\right)^4`$ (2)
$`+1.66\times 10^{12}`$
where the $`1.66\times 10^{12}`$ comes from small hadronic and weak loop effects. Assuming no significant โnew physicsโ contributions to $`a_e^{\mathrm{th}}`$, it can be equated with (1) to give
$$\alpha ^1=137.03599959(40)$$
(3)
That precision is already very impressive. Improvement by a factor of 10 appears to be technically feasible gabrielse and should certainly be undertaken. However, at this time such improvement would not further our ability to test QED. Pure QED tests require comparable measurements of $`\alpha `$ in other processes. Agreement between two distinct $`\alpha `$ determinations tests QED and probes for โnew physicsโ effects. After $`a_e`$, the next best (direct) measurement of $`\alpha `$ comes from the quantum Hall effect
$$\alpha ^1(qH)=137.03600370(270)$$
(4)
which is not nearly as precise. Nevertheless, the agreement of (3) and (4) (at the 1.50 sigma level) is a major triumph for QED up to the 4 loop quantum level.
In terms of probing โnew physicsโ, one can search for a shift in $`a_e`$ by $`m_e^2/\mathrm{\Lambda }_e^2`$ where $`\mathrm{\Lambda }_e`$ is the approximate scale of some generic new short-distance effect. Current comparison of $`a_e\alpha `$ and $`\alpha (qH)`$ explores $`\mathrm{\Lambda }_e<\mathrm{\hspace{0.33em}100}`$ GeV. To probe the much more interesting $`\mathrm{\Lambda }_e๐ช`$ (TeV) region would require an order of magnitude improvement in $`a_e`$ and about two orders of magnitude error reduction in some direct precision determination of $`\alpha `$ such as the quantum Hall effect. Perhaps the most likely possibility is to use the already very precisely measured Rydberg constant in conjunction with a much improved $`m_e`$ determination to obtain an independent $`\alpha `$.
The usual fine structure constant, $`\alpha `$, is defined at zero momentum transfer as is appropriate for low energy atomic physics phenomena. However, that definition is not well suited for short-distance electroweak effects. Vacuum polarization loops screen charges such that the effective (running) electric charge increases at short-distances. One can incorporate those quantum loop contributions into a short-distance prd20274 $`\alpha (m_Z)`$ defined at $`q^2=m_Z^2`$. The main effect comes from lepton loops, which can be very precisely calculated, and somewhat smaller hadronic loops. The latter are not as theoretically clean and must be obtained by combining perturbative calculations with results of a dispersion relation which employs $`๐ช(e^+e^{}`$hadrons) data. A detailed study by Davier and Hรถcker found plb439427
$$\alpha ^1(m_Z)=128.933(21)$$
(5)
where the uncertainty stems primarily from low energy hadronic loops. Although not nearly as precise as $`\alpha ^1`$, the uncertainty quoted in (5) is impressively small and a tribute to the effort that has gone into reducing it. (When I first studied this issue in 1979, I crudely estimated prd20274 $`\alpha ^1(m_Z)128.5\pm 1.0`$.) However, the error in (5) is still somewhat controversial, primarily because of its reliance on perturbative QCD down to very low energies. For comparison, an earlier study by Eidelman and Jegerlehner zpc67585 , which relied less on perturbative QCD and more on $`e^+e^{}`$ data found
$$\alpha ^1(m_Z)=128.896(90)(\mathrm{E}\&\mathrm{J}1995)$$
(6)
That estimated uncertainty is often cited as more conservative and therefore sometimes employed in $`m_H`$ and โnew physicsโ constraints. As we shall see, the smaller uncertainty in (5) has very important consequences for predicting the Higgs mass. I note that a more recent study jeger by Eidelman and Jegerlehner finds
$$\alpha ^1(m_Z)=128.913(35)(\mathrm{E}\&\mathrm{J}1998)$$
(7)
which is in good accord with (5) and also exhibits relatively small uncertainty. In my subsequent discussion, I employ the result in (5), but caution the reader that a more conservative approach would expand the uncertainty, perhaps even by as much as a factor of 4.
A related short-distance coupling, $`\alpha (m_Z)_{\overline{MS}}`$, can be defined by modified minimal subtraction at scale $`\mu =m_Z`$. It is particularly useful for studies of coupling unification in grand unified theories (GUTS) where a uniform comparitive definition $`(\overline{MS})`$ of all couplings is called for prl46163 . The quantities $`\alpha (m_Z)`$ and $`\alpha (m_Z)_{\overline{MS}}`$ differ by a constant, such that hepph9803453
$$\alpha ^1(m_Z)_{\overline{MS}}=\alpha ^1(m_Z)0.982=127.951(21)$$
(8)
In weak interaction physics, the most precisely determined parameter is the Fermi constant, $`G_\mu `$, as obtained from the muon lifetime. One extracts that quantity by comparing the experimental value
$$\tau _\mu =2.197035(40)\times 10^6s$$
(9)
with the theoretical prediction
$`\tau _\mu ^1=\mathrm{\Gamma }(\mu \mathrm{all})`$ $`=`$ $`{\displaystyle \frac{G_\mu ^2m_\mu ^5}{192\pi ^3}}f\left({\displaystyle \frac{m_e^2}{m_\mu ^2}}\right)(1+\mathrm{R}.\mathrm{C}.)(1+{\displaystyle \frac{3}{5}}{\displaystyle \frac{m_\mu ^2}{m_W^2}})`$
$`f(x)`$ $`=`$ $`18x+8x^3x^412x^2\mathrm{}nx`$ (10)
In that expression R.C. stands for Radiative Corrections. Those terms are somewhat arbitrary in the standard model. The point being that $`G_\mu `$ is a renormalized parameter which is used to absorb most loop corrections to muon decay. Those corrections not absorbed into $`G_\mu `$ are explicitly factored out in R.C. For historical reasons and in the spirit of effective field theory approaches, R.C. has been chosen to be the QED corrections to the old V-A four fermion description of muon decay ap2020 . That definition is practical, since the QED corrections to muon decay in the old V-A theory are finite to all orders in perturbation theory. In that way, one finds
$$\mathrm{R}.\mathrm{C}.=\frac{\alpha }{2\pi }(\frac{25}{4}\pi ^2)(1+\frac{\alpha }{\pi }(\frac{2}{3}\mathrm{}n\frac{m_\mu }{m_e}3.7)+\left(\frac{\alpha }{\pi }\right)^2(\frac{4}{9}\mathrm{}n^2\frac{m_\mu }{m_e}2.0\mathrm{}n\frac{m_\mu }{m_e}+C)\mathrm{})$$
(11)
The leading $`๐ช(\alpha )`$ terms in that expression have been known for a long time from the pioneering work of Kinoshita and Sirlin pr1131652 and Berman pr112267 . Coefficients of the higher order logs can be obtained from the renormalization group constraint npb29296
$`(m_e{\displaystyle \frac{}{m_e}}+\beta (\alpha ){\displaystyle \frac{}{\alpha }})\mathrm{R}.\mathrm{C}.=0`$
$`\beta (\alpha )={\displaystyle \frac{2}{3}}{\displaystyle \frac{\alpha ^2}{\pi }}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{\alpha ^3}{\pi ^2}}\mathrm{}`$ (12)
The -3.7 two loop constant in parenthesis was recently computed by van Ritbergen and Stuart prl82488 . It almost exactly cancels the leading log two loop correction obtained from the renormalization group approach (or mass singularities argument) of Roos and Sirlin npb29296 . Hence, the original $`๐ช(\alpha )`$ correction in (9) is a much better approximation than one might have guessed. Comparing (9) and (10), one finds
$$G_\mu =1.16637(1)\times 10^5\mathrm{GeV}^2$$
(13)
There have been several experimental proposals czar to reduce the uncertainty in $`\tau _\mu `$ and $`G_\mu `$ by as much as a factor of 20. Such improvement appears technically feasible and, given the fundamental nature of $`G_\mu `$, should certainly be undertaken. However, from the point of view of testing the standard model, the situation is similar to $`\alpha `$. $`G_\mu `$ is already much better known than the other parameters it can be compared with; so, significant improvement must be made in other quantities before a more precise $`G_\mu `$ is required.
Let me emphasize the fact that lots of interesting loop effects have been absorbed into the renormalization of $`g_{2_0}^2/4\sqrt{2}m_W^{0^2}`$ which we call $`G_\mu `$. Included are top quark npb12389 and Higgs loop corrections npb84132 to the $`W`$ boson propagator as well as potential โnew physicsโ from SUSY loops, Technicolor etc. Even tree level effects of possible more massive gauge bosons such as excited $`W^^\pm `$ bosons could be effectively incorporated into $`G_\mu `$. To uncover those contributions requires comparison of $`G_\mu `$ with other precisely measured electroweak parameters which have different quantum loop (or tree level) dependences. Of course, those quantities must be related to $`G_\mu `$ in such a way that short-distance divergences cancel in the comparison.
Fortunately, due to an underlying global SU(2)<sub>V</sub> symmetry in the standard model, there exist natural relations among various bare parameters nc16a423
$$\mathrm{sin}^2\theta _W^0=\frac{e_0^2}{g_{2_0}^2}=1(m_W^0/m_Z^0)^2$$
(14)
Each of those bare unrenormalized expressions contains short-distance infinities; however, because the theory is renormalizable, the divergences are the same. Therefore, those relations continue to hold for renormalized quantities, up to finite, calculable radiative corrections nc16a423 . The residual radiative corrections contain very interesting effects such as $`m_t`$ and $`m_H`$ dependence as well as possible โnew physicsโ. So, for example, one can relate
$$G_\mu =\frac{\pi \alpha }{\sqrt{2}m_W^2(1m_W^2/m_Z^2)}(1+rad.corr.)$$
(15)
and test the predicted radiative corrections, if $`m_Z`$ and $`m_W`$ are also precisely known.
Gauge boson masses are not as well determined as $`G_\mu `$, but they have reached high levels of precision. In particular, the $`Z`$ mass has been measured with high statistics Breit-Wigner fits to the $`Z`$ resonance at LEP with the result swartz
$$m_Z=91.1871(21)\mathrm{GeV}$$
(16)
That determination is so good that one must be very precise regarding the definition of $`m_Z`$. (Remember the $`Z`$ has a relatively large width $`2.5`$ GeV.) The quantity in (16) is related to the real part of the $`Z`$ propagator pole, $`m_Z`$ (pole), and full width, $`\mathrm{\Gamma }_Z`$, by prl672127
$$m_Z^2=m_Z^2(\mathrm{pole})+\mathrm{\Gamma }_Z^2$$
(17)
The two mass definitions $`m_Z`$ and $`m_Z`$ (pole) differ by about 34 MeV, which is much larger than the uncertainty in (16). Hence, one must specify which definition is being employed in precision studies. I note, that the $`m_Z`$ in (16) is also more appropriate for use in low energy neutral current amplitudes.
In the case of the $`W^\pm `$ bosons, the renormalized mass, $`m_W`$, is similarly defined by
$$m_W^2=m_W^2(\mathrm{pole})+\mathrm{\Gamma }_W^2$$
(18)
That quantity is obtained lanc from studies at $`p\overline{p}`$ colliders, $`m_W=80.448(62)`$ GeV, as well as $`e^+e^{}W^+W^{}`$ at LEPII, $`m_W=80.401(48)`$ GeV. Together they average to
$$m_W=80.419(38)\mathrm{GeV}.$$
(19)
The current level of uncertainty, $`\pm 38`$ MeV, is large compared to $`\mathrm{\Delta }m_Z`$. It is expected that continuing efforts at LEPII and Run II at Fermilabโs Tevatron should reduce that error to about $`\pm 25`$ MeV. A challenging but worthwhile goal for future high energy facilities would be to push $`\mathrm{\Delta }m_W`$ to $`\pm 10`$ MeV or better. At that level, all sorts of interesting โnew physicsโ effects are probed. I note that the $`m_W`$ defined in (19) is also the appropriate quantity for low energy amplitudes such as muon decay.
Another important quantity for precision standard model tests is $`m_t`$, the top quark mass. Measurements from CDF and D$`\mathrm{}`$ at Fermilab give lanc
$$m_t(\mathrm{pole})=174.3\pm 5.1\mathrm{GeV}$$
(20)
Reducing that uncertainty further is important for quantum loop studies, as we shall subsequently see. Future Tevatron Run II efforts are expected to reduce the error in $`m_t`$ to about $`\pm 2`$ GeV. LHC and NLC studies should bring it well below $`\pm 1`$ GeV.
In addition to masses, the renormalized weak mixing angle plays a central role in tests of the standard model. That parameter can be defined in a variety of ways, each of which has its own advocates. I list three popular examples prd20274 ; pr491160 ; czartwo
$`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ $`(\overline{MS}\mathrm{definition}\mathrm{at}\mu =m_Z)(a)`$
$`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ $`(Z\mu \overline{\mu }\mathrm{vertex})(b)`$
$`\mathrm{sin}^2\theta _W`$ $`1m_W^2/m_Z^2(c)`$
They differ by finite $`๐ช(\alpha )`$ loop corrections. The $`\overline{MS}`$ definition is particularly simple, being defined as the ratio of two $`\overline{MS}`$ couplings $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}e^2(m_Z)_{\overline{MS}}/`$ $`g_2^2(m_Z)_{\overline{MS}}`$. It was introduced for GUT studies prl46163 , but is useful for most electroweak analyses. The effective, $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$, weak angle was invented for $`Z`$ pole analyses. Roughly speaking, it is defined by the ratio of vector and axial-vector components (including loops) for the on-mass-shell $`Z\mu \overline{\mu }`$ vertex$`1`$$`4\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$. Although conceptually rather simple, analytic electroweak radiative corrections expressed in terms of $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ are complicated and ugly. Numerically, it is close to the $`\overline{MS}`$ definition pr491160
$$\mathrm{sin}^2\theta _W^{\mathrm{eff}}=\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}+0.00028$$
(22)
but the analytic structure of the difference is quite complicated. For those intent on employing $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$, a strategy might be to calculate radiative corrections in terms of $`\mathrm{sin}^2\theta _W`$ $`(m_Z)_{\overline{MS}}`$ and then translate to $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ via (22). But why not simply use $`\mathrm{sin}^2\theta _W`$ $`(m_Z)_{\overline{MS}}`$?
Currently, $`Z`$ pole studies at LEP and SLAC give swartz
$`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ $`=`$ $`0.23091\pm 0.00021`$
$`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ $`=`$ $`0.23119\pm 0.00021`$ (23)
That result includes measurements of the left-right asymmetry, $`A_{LR}`$, at SLAC as well as the various lepton asymmetries at LEP and SLAC. The $`A_{LR}`$ contribution has for some time given a relatively low value for the weak mixing angle. The latest swartz SLD result is
$$\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}=0.23073(28)$$
(24)
Currently, the $`Zb\overline{b}`$ forward-backward asymmetry at LEP gives a higher $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ and, if included, would bring up the average. However, the $`Zb\overline{b}`$ coupling appears to be somewhat anomalous and suggests problematic $`b`$ identification; so, one should be cautious when including such results in averages.
There are very good reasons to clarify and further improve $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$. One could imagine redoing $`A_{LR}`$ at a future polarized lepton-lepton ($`e^+e^{}`$ or $`\mu ^+\mu ^{}`$) collider, but with much higher statistics. In principle, one might reduce the uncertainty in $`\mathrm{sin}^2\theta _W^{\mathrm{eff}}`$ by a factor of 10 to $`\pm 0.00002`$, an incredible achievement if accomplished czartwo .
The so-called on-shell or mass definition prd22971 in (Fundamental Parameters and Precision Measurementsc) also has its advocates. It can be directly obtained from $`m_W`$ and $`m_Z`$ determinations. Indeed, at hadron colliders, the ratio $`m_W/m_Z`$ can have reduced systematic uncertainties. One could imagine that the current uncertainty in
$$\mathrm{sin}^2\theta _W=1m_W^2/m_Z^2=0.2222\pm 0.0007$$
(25)
might be reduced by a factor of about 4 at the LHC. Such a reduction is extremely important since the comparison of $`\mathrm{sin}^2\theta _W`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ provides a clean probe of โnew physicsโ. It is also possible (because of a subtle cancellation of certain loop effects npb189442 ) to measure $`\mathrm{sin}^2\theta _W`$ more directly in deep-inelastic $`\nu _\mu N`$ scattering. Indeed, a recent Fermilab experiment found epjc1509
$$\mathrm{sin}^2\theta _W=0.2255\pm 0.0019\pm 0.0010$$
(26)
where the first error is statistical and the second systematic. That single measurement is quite competitive with (25) and complements it nicely. One might imagine a future high statistics effort significantly reducing the error in (26), but that would require a new, intense, high energy neutrino beam.
Two other well measured electroweak parameters are the charged and neutral leptonic partial widths of the $`Z`$ boson swartz
$`\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))`$ $`=`$ $`83.96\pm 0.09\mathrm{MeV}`$
$`\mathrm{\Gamma }(Z\mathrm{\Sigma }\nu \overline{\nu })`$ $`=`$ $`498.8\pm 1.5\mathrm{MeV}`$ (27)
The first of those, by definition, corresponds to $`Z`$ decay into massless charged leptons along with the possibility of inclusive bremsstrahlung. The second represents the inclusive invisible width of the $`Z`$.
All of the above precision measurements can be collectively used to test the standard model, predict the Higgs mass, and search for โnew physicsโ effects. That ability stems from the natural relations in (14) and calculations prd22971 ; prd222695 of the radiative corrections to them. Parametrizing those radiative corrections by $`\mathrm{\Delta }r`$, $`\mathrm{\Delta }r(m_Z)_{\overline{MS}}`$, and $`\mathrm{\Delta }\widehat{r}`$, one finds npb35149
$`{\displaystyle \frac{\pi \alpha }{\sqrt{2}G_\mu m_W^2}}`$ $`=`$ $`\left(1{\displaystyle \frac{m_W^2}{m_Z^2}}\right)(1\mathrm{\Delta }r)(a)`$
$`{\displaystyle \frac{\pi \alpha }{\sqrt{2}G_\mu m_W^2}}`$ $`=`$ $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}(1\mathrm{\Delta }r(m_Z)_{\overline{MS}})(b)`$ (28)
$`{\displaystyle \frac{4\pi \alpha }{\sqrt{2}G_\mu m_Z^2}}`$ $`=`$ $`\mathrm{sin}^22\theta _W(m_Z)_{\overline{MS}}(1\mathrm{\Delta }\widehat{r})(c)`$
Those expressions contain all one loop corrections to $`\alpha `$, muon decay, $`m_W`$, $`m_Z`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ and incorporate some leading two loop contributions. The quantities $`\mathrm{\Delta }r`$ and $`\mathrm{\Delta }\widehat{r}`$ are particularly interesting because of their dependence on $`m_t`$ and $`m_H`$. In addition, all three quantities provide probes of โnew physicsโ.
Numerically, all three radiative corrections in (28) contain a significant contribution from vacuum polarization effects prd20274 in $`\alpha `$, about $`+7\%`$. They are basically the same as the corrections that enter into the evolution of $`\alpha `$ to $`\alpha (m_Z)`$. Leptonic loops contribute a significant part of that effect and can be very accurately computed. Hadronic loops are less clean theoretically and lead to a common uncertainty in $`\mathrm{\Delta }r`$, $`\mathrm{\Delta }r(m_Z)_{\overline{MS}}`$, and $`\mathrm{\Delta }\widehat{r}`$ of
$$\alpha \mathrm{\Delta }\alpha ^1(m_Z)$$
(29)
For $`\mathrm{\Delta }\alpha ^1(m_Z)=0.021`$ as in (5), that amounts to a rather negligible $`\pm 0.00015`$ error. However, for $`\mathrm{\Delta }\alpha ^1(m_Z)=\pm 0.090`$ as in (6), it increases to $`\pm 0.00066`$. That large an uncertainty would impact precision tests. If one wishes to avoid that low energy hadronic loop uncertainty, dependence on $`\alpha `$ can be circumvented by considering
$$\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}=\left(1\frac{m_W^2}{m_Z^2}\right)(1\mathrm{\Delta }r+\mathrm{\Delta }r(m_Z)_{\overline{MS}})$$
(30)
Currently, that comparison suggests a very light Higgs, but is not yet quite competitive with eq. (28c) in constraining $`m_H`$. However, future improvements in $`m_W`$ and $`m_t`$ could make it very interesting.
Another useful relation that provides some sensitivity to $`m_H`$ without $`\mathrm{\Delta }\alpha `$ uncertainties involves the partial $`Z`$ width when written as marc
$$\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))=\frac{G_\mu m_Z^3(14\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}+8\mathrm{sin}^4\theta _W(m_Z)_{\overline{MS}})}{12\sqrt{2}\pi (1\mathrm{\Delta }r_Z(m_H))}$$
(31)
Using $`m_t=174.3\pm 5.1`$ GeV as input, one can compute the radiative corrections in (28) as functions of $`m_H`$. Those results are illustrated in table 1. Note that $`\mathrm{\Delta }r`$ is most sensitive to changes in $`m_H`$ but also carries the largest uncertainty from $`\mathrm{\Delta }m_t=\pm 5.1`$ GeV ($`\pm 0.0020`$). Hence, efforts to determine $`m_H`$ from $`m_W`$ are starting to require a better measurement of $`m_t`$. On the other hand, determining $`m_H`$ from $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ via $`\mathrm{\Delta }\widehat{r}`$ is less sensitive to $`\mathrm{\Delta }m_t`$ but more sensitive to $`\mathrm{\Delta }\alpha ^1(m_Z)`$. Those dependences as well as $`\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))`$ are illustrated by the following approximate relations plb418188 which are very accurate up to $`m_H๐ช`$(400 GeV)
$`m_W`$ $`=`$ $`(80.385\pm 0.032\pm 0.003\mathrm{GeV})(10.00072\mathrm{}n\left({\displaystyle \frac{m_H}{100\mathrm{GeV}}}\right)`$ (32)
$`1\times 10^4\mathrm{}n^2\left({\displaystyle \frac{m_H}{100\mathrm{GeV}}}\right))`$
$`\mathrm{sin}^2\theta _W\left(m_Z\right)_{\overline{MS}}`$ $`=`$ $`\left(0.23112\pm 0.00016\pm 0.00006\right)\left(1+0.00226\mathrm{}n\left({\displaystyle \frac{m_H}{100\mathrm{GeV}}}\right)\right)`$
$`\mathrm{\Gamma }\left(Z\mathrm{}^+\mathrm{}^{}\left(\gamma \right)\right)`$ $`=`$ $`(84.011\pm 0.047\pm 0.0028\mathrm{MeV})(10.00064\mathrm{}n\left({\displaystyle \frac{m_H}{100\mathrm{GeV}}}\right)`$
$`0.00026\mathrm{}n^2\left({\displaystyle \frac{m_H}{100\mathrm{GeV}}}\right)`$
where the errors correspond to $`\mathrm{\Delta }m_t=\pm 5.1`$ GeV and $`\mathrm{\Delta }\alpha ^1(m_Z)=\pm 0.021`$ respectively. Note that increasing $`\mathrm{\Delta }\alpha ^1(m_Z)`$ to $`\pm 0.090`$ would significantly compromise the utility of $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ for determining $`m_H`$ but have less of an impact on $`m_W`$. Predictions for $`m_W`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ are illustrated in table 2 for various $`m_H`$ values plb394188 .
Employing $`m_W=80.419(38)`$ GeV, $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}=0.23091(21)`$, and $`\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))=83.96(9)`$ MeV one finds from eq (32)
$`m_H`$ $`=`$ $`53_{40}^{+77}\mathrm{GeV}(\mathrm{from}m_W)`$ (33)
$`m_H`$ $`=`$ $`67_{27}^{+45}\mathrm{GeV}(\mathrm{from}\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}})`$ (34)
$`m_H`$ $`=`$ $`208_{180}^{+340}\mathrm{GeV}(\mathrm{from}\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))`$ (35)
Several features of those predictions are revealing. The first is that $`\mathrm{sin}^2\theta _W`$ $`(m_Z)_{\overline{MS}}`$ currently gives the best determination of $`m_H`$. Note, however, the uncertainties scale as the central value; so, the relatively small value, 67 GeV, helps reduce the uncertainties. Also, a larger $`\mathrm{\Delta }\alpha ^1(m_Z)=\pm 0.090`$ would significantly increase the overall uncertainty hepph9812332 . In the case of $`m_W`$, a light Higgs is also suggested. In fact, the $`m_W`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ determinations of $`m_H`$ are very consistent.
Taken together, (33) and (34) appear to point to a relatively light Higgs scalar that cannot be too far from the current LEPII bound based on the non observation of $`e^+e^{}ZH`$ wu
$$m_H>106\mathrm{GeV}$$
(36)
In the near future, searching for the Higgs via associated $`W^\pm H`$ and $`ZH`$ at the Fermilab $`p\overline{p}`$ collider during Run II promises discovery up to $`m_H115`$โ130 GeV, perhaps even higher. Higgs unveiling may soon be at hand.
## Discussion
Within the Standard Model framework, precision electroweak studies suggest a relatively light Higgs scalar. The upper bound from global fits to all data is about erler 235 GeV. Such a bound is in accord with โno new physicsโ scenarios, where perturbative validity up to $`๐ช(10^{19}`$ GeV) and vacuum stability imply the range sirlin
$$135\mathrm{GeV}<m_H<\mathrm{\hspace{0.33em}180}\mathrm{GeV}$$
(37)
Supersymmetry, on the other hand, is starting to be squeezed by lack of a Higgs scalar discovery at LEP II. The MSSM scenario requires $`m_H<\mathrm{\hspace{0.33em}135}`$ GeV, but one generally finds that considerably lower values are preferred. The current bound of $`m_H>\mathrm{\hspace{0.33em}106}`$ GeV will be pushed to $`112`$ GeV at LEP II and somewhat further by Run II at the Fermilab Tevatron. If MSSM has some relation to Natureโs truth, a discovery should not be far off.
What about alternative theories where there is no fundamental Higgs? In some dynamical electroweak symmetry breaking scenarios, one expects an effective $`m_H๐ช`$(1 TeV). Are such ideas ruled out? They face a basic difficulty which is nicely illustrated by the $`S`$ & $`T`$ parameters of Peskin and Takeuchi peskin . For $`m_H1`$ TeV, one finds the following predictions
$`m_W`$ $``$ $`80.214\mathrm{GeV}(10.0036S+0.0056T)`$
$`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ $``$ $`0.23250(1+0.0158S0.0113T)`$ (38)
$`\mathrm{\Gamma }(Z\mathrm{}^+\mathrm{}^{}(\gamma ))`$ $``$ $`83.79\mathrm{MeV}(10.0021S+0.0093T)`$
where $`S`$ and $`T`$ represent loop effects from heavy, strongly interacting, condensate fermions. Comparison of $`m_W`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ via marctwo
$$S118\left\{2\left(\frac{m_W80.214\mathrm{GeV}}{80.214\mathrm{GeV}}\right)+\frac{\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}0.23250}{0.23250}\right\}$$
(39)
suggests from $`m_W=80.419(38)`$ GeV and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}=0.23091(21)`$
$$S0.20\pm 0.15$$
(40)
while consistency with eq. (38) requires
$$T0.33\pm 0.17$$
(41)
Generating $`T0.3`$ by loop effects is actually quite easy. It requires relatively small mass splittings of new heavy fermion isodoublets. The problem is a negative $`S`$ which is possible, but not particularly natural in simple dynamical models where $`S๐ช(1)`$ is more likely. However, one could easily imagine small shifts in $`m_W`$ and $`\mathrm{sin}^2\theta _W(m_Z)_{\overline{MS}}`$ which could give $`S0`$ and $`T>0`$. So, $`m_H1`$ TeV could be accommodated by non-vanishing $`S`$ & $`T`$, but it is much easier to simply satisfy the precision measurement constraints with $`m_H100`$โ200 GeV and $`ST0`$. |
warning/0003/physics0003037.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Among the numerous and important questions offered to the physicist by the sciences of life, the ones relative to the genetic code present a particular interest. Indeed, in addition to the fundamental importance of this domain, the DNA structure on the one hand and the mechanism of polypeptid fixation from codons on the other hand possess appealing aspects for the theorist. Let us, in a brief summary, select some essential features . First, as well known, the DNA macromolecule is constituted by two linear chains of nucleotides in a double helix shape. There are four different nucleotides, characterized by their bases: adenine (A) and guanine (G) deriving from purine, and cytosine (C) and thymine (T) coming from pyrimidine. Note also that an A (resp. T) base in one strand is connected with two hydrogen bonds to a T (resp. A) base in the other strand, while a C (resp. G) base is related to a G (resp. C) base with three hydrogen bonds. The genetic information is transmitted to the cytoplasm via the messenger ribonucleic acid or mRNA. During this operation, called transcription, the A, G, C, T bases in the DNA are associated respectively to the U, C, G, A bases, U denoting the uracile base. Then it will be through a ribosome that a triplet of nucleotides or codon will be related to an amino-acid. More precisely, a codon is defined as an ordered sequence of three nucleotides, e.g. AAG, ACG, etc., and one enumerates in this way $`4\times 4\times 4=64`$ different codons. Following the universal eukariotic code (see Table 4), 61 of such triplets can be connected in an unambiguous way to the amino-acids, except the three following triplets UAA, UAG and UGA, which are called non-sense or stop-codons, the role of which is to stop the biosynthesis. Indeed, the genetic code is the association between codons and amino-acids. But since one distinguishes only 20 amino-acids<sup>1</sup><sup>1</sup>1Alanine (Ala), Arginine (Arg), Asparagine (Asn), Aspartic acid (Asp), Cysteine (Cys), Glutamine (Gln), Glutamic acid (Glu), Glycine (Gly), Histidine (His), Isoleucine (Ile), Leucine (Leu), Lysine (Lys), Methionine (Met), Phenylalanine (Phe), Proline (Pro), Serine (Ser), Threonine (Thr), Tryptophane (Trp), Tyrosine (Tyr), Valine (Val). related to the 61 codons, it follows that the genetic code is degenerated. Still considering the standard eukariotic code, one observes sextets, quadruplets, triplets, doublets and singlets of codons, each multiplet corresponding to a specific amino-acid. Such a picture naturally suggests to look for an underlying symmetry able to describe the observed structure in multiplets, in the spirit of dynamical symmetry scheme which has proven so powerful in atomic, molecular and nuclear physics. We review at the end of this paper these recent approaches.
In refs. we have proposed a mathematical framework in which the codons appear as composite states of nucleotides. The four nucleotides being assigned to the fundamental irreducible representation of the quantum group $`๐ฐ_q(sl(2)sl(2))`$ in the limit $`q0`$, the codons are obtained as tensor product of nucleotides. Indeed, the properties of quantum group representations in the limit $`q0`$, or crystal basis, are well adapted to take into account the nucleotide ordering. Then properties of this model have been considered. We will generalize some of them in the following and also propose new developments.
The paper is organized as follows. We start in sect. 2 by recalling the main aspects of the model. In sect. 3 we build out of the generators of $`๐ฐ_{q0}(sl(2)sl(2))`$ a reading operator, which gives the correct correspondence between codons and amino-acids for each of the 12 presently known genetic codes. This construction generalizes in a synthetical way the one started in for the eukariotic and vertebrate mitochondrial codes, the different reading operators acting on codons and providing the same eigenvalue for a given amino-acid whatever the considered code. In sect. 4 some physical properties of dinucleotide states are fitted. In sect. 5, we analyze ratios of codon usage frequency for several biological species belonging to the vertebrate class and put in evidence a universal behaviour, which fits naturally in our model. In sect. 6, making use of the general crystal basis mathematical framework, we represent the mutation induced by the deletion of a pyrimidine by the action of a suitable crystal spinor operator. In sect. 7 we review and compare with our model the recent symmetry approaches to the genetic code. Finally in sect. 8 we give a few conclusions and discuss some directions of future developments.
## 2 The Model
We consider the four nucleotides as basic states of the $`(\frac{1}{2},\frac{1}{2})`$ representation of the $`๐ฐ_q(sl(2)sl(2))`$ quantum enveloping algebra in the limit $`q0`$. A triplet of nucleotides will then be obtained by constructing the tensor product of three such four-dimensional representations. Actually, this approach mimicks the group theoretical classification of baryons made out from three quarks in elementary particles physics, the building blocks being here the A, C, G, T/U nucleotides. The main and essential difference stands in the property of a codon to be an *ordered* set of three nucleotides, which is not the case for a baryon.
Constructing such pure states is made possible in the framework of any algebra $`๐ฐ_{q0}(๐ข)`$ with $`๐ข`$ being any (semi)-simple classical Lie algebra owing to the existence of a special basis, called crystal basis, in any (finite dimensional) representation of $`๐ข`$. The algebra $`๐ข=sl(2)sl(2)`$ appears the most natural for our purpose. The complementary rule in the DNAโmRNA transcription may suggest to assign a *quantum number* with opposite values to the couples (A,T/U) and (C,G). The distinction between the purine bases (A,G) and the pyrimidine ones (C,T/U) can be algebraically represented in an analogous way. Thus considering the fundamental representation $`(\frac{1}{2},\frac{1}{2})`$ of $`sl(2)sl(2)`$ and denoting $`\pm `$ the basis vector corresponding to the eigenvalues $`\pm \frac{1}{2}`$ of the $`J_3`$ generator in any of the two $`sl(2)`$ corresponding algebras, we will assume the following โbiologicalโ spin structure:
$`sl(2)_H`$
$`C(+,+)`$ $``$ $`U(,+)`$
$`.sl(2)_V`$ $`sl(2)_V`$ (1)
$`.G(+,)`$ $``$ $`A(,)`$
$`sl(2)_H`$
the subscripts $`H`$ (:= horizontal) and $`V`$ (:= vertical) being just added to specify the algebra.
Now, we consider the representations of $`๐ฐ_q(sl(2))`$ and more specifically the crystal bases obtained when $`q0`$. Introducing in $`๐ฐ_{q0}(sl(2))`$ the operators $`J_+`$ and $`J_{}`$ after modification of the corresponding simple root vectors of $`๐ฐ_q(sl(2))`$, a particular kind of basis in a $`๐ฐ_q(sl(2))`$-module can be defined. Such a basis is called a crystal basis and carries the property to undergo in a specially simple way the action of the $`J_+`$ and $`J_{}`$ operators: as an example, for any couple of vectors $`u,v`$ in the crystal basis $``$, one gets $`u=J_+v`$ if and only if $`v=J_{}u`$. More interesting for our purpose is the crystal basis in the tensorial product of two representations. Then the following theorem holds (written here in the case of $`sl(2)`$):
###### Theorem 1 (Kashiwara)
Let $`_1`$ and $`_2`$ be the crystal bases of the $`M_1`$ and $`M_2`$ $`๐ฐ_{q0}(sl(2))`$-modules respectively. Then for $`u_1`$ and $`v_2`$, we have:
$`J_{}(uv)=\{\begin{array}{cc}J_{}uv\hfill & n1\text{ such that }J_{}^nu0\text{ and }J_+v=0\hfill \\ uJ_{}v\hfill & \text{otherwise}\hfill \end{array}`$ (5)
$`J_+(uv)=\{\begin{array}{cc}uJ_+v\hfill & n1\text{ such that }J_+^nv0\text{ and }J_{}u=0\hfill \\ J_+uv\hfill & \text{otherwise}\hfill \end{array}`$ (8)
Note that the tensor product of two representations in the crystal basis is not commutative. However, in the case of our model, we only need to construct the $`n`$-fold tensor product of the fundamental representation $`(\frac{1}{2},\frac{1}{2})`$ of $`๐ฐ_{q0}(sl(2)sl(2))`$ by itself, thus preserving commutativity and associativity.
Let us insist on the choice of the crystal basis, which exists only in the limit $`q0`$. In a codon the order of the nucleotides is of fundamental importance (e.g. CCU $``$ Pro, CUC $``$ Leu, UCC $``$ Ser). If we want to consider the codons as composite states of the (elementary) nucleotides, this surely cannot be done in the framework of Lie (super)algebras. Indeed in the Lie theory, the composite states are obtained by performing tensor products of the fundamental irreducible representations. They appear as linear combinations of the elementary states, with symmetry properties determined from the tensor product (i.e. for $`sl(n)`$, by the structure of the corresponding Young tableaux). On the contrary the crystal basis provides us with the mathematical structure to build composite states as *pure* states, characterized by the order of the constituents. In order to dispose of such a basis, we need to consider the limit $`q0`$. Note that in this limit we do not deal anymore either with a Lie algebra or with an universal deformed enveloping algebra.
To represent a codon, we have to perform the tensor product of three $`(\frac{1}{2},\frac{1}{2})`$ representations of $`๐ฐ_{q0}(sl(2)sl(2))`$. However, it is well-known (see Tables 4) that in a multiplet of codons relative to a specific amino-acid, the two first bases constituent of a codon are โrelatively stableโ, the degeneracy being mainly generated by the third nucleotide. We consider first the tensor product:
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})=(1,1)(1,0)(0,1)(0,0)$$
(9)
where inside the parenthesis, $`j=0,\frac{1}{2},1`$ is put in place of the $`2j+1=1,2,3`$ respectively dimensional $`sl(2)`$ representation. We get, using Theorem 1, the following tableau:
$$\begin{array}{ccccc}su(2)_H\hfill & (0,0)& (\text{CA})& (1,0)& (\begin{array}{ccc}\text{CG}& \text{UG}& \text{UA}\end{array})\\ \hfill & & & & \\ su(2)_V\hfill & (0,1)& \left(\begin{array}{c}\text{CU}\\ \text{GU}\\ \text{GA}\end{array}\right)& (1,1)& \left(\begin{array}{ccc}\text{CC}& \text{UC}& \text{UU}\\ \text{GC}& \text{AC}& \text{AU}\\ \text{GG}& \text{AG}& \text{AA}\end{array}\right)\end{array}$$
From Table 4, the dinucleotide states formed by the first two nucleotides in a codon can be put in correspondence with quadruplets, doublets or singlets of codons relative to an amino-acid. Note that the sextets (resp. triplets) are viewed as the sum of a quadruplet and a doublet (resp. a doublet and a singlet). Let us define the โchargeโ $`Q`$ of a dinucleotide state by
$$Q=J_{H,3}^{(1)}+J_{H,3}^{(2)}+J_{V,3}^{(2)}$$
(10)
where the superscript $`(1)`$ or $`(2)`$ denotes the position of a codon in the dinucleotide state.
The dinucleotide states are then split into two octets with respect to the charge $`Q`$: the eight *strong* dinucleotides associated to the quadruplets (as well as those included in the sextets) of codons satisfy $`Q>0`$, while the eight *weak* dinucleotides associated to the doublets (as well as those included in the triplets) and eventually to the singlets of codons satisfy $`Q<0`$. Let us remark that by the change $`CA`$ and $`UG`$, which is equivalent to the change of the sign of $`J_{3,\alpha }`$ or to reflexion with respect to the diagonals of the eq.(LABEL:eq:gc1), the 8 strong dinucleotides are transformed into weak ones and vice-versa.
If we consider the three-fold tensor product, the content into irreducible representations of $`๐ฐ_{q0}(sl(2)sl(2))`$ is given by:
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})=(\frac{3}{2},\frac{3}{2})2(\frac{3}{2},\frac{1}{2})2(\frac{1}{2},\frac{3}{2})4(\frac{1}{2},\frac{1}{2})$$
(11)
The structure of the irreducible representations of the r.h.s. of Eq. (11) is (the upper labels denote different irreducible representations):
$`(\frac{3}{2},\frac{3}{2})\left(\begin{array}{cccc}\text{CCC}& \text{UCC}& \text{UUC}& \text{UUU}\\ \text{GCC}& \text{ACC}& \text{AUC}& \text{AUU}\\ \text{GGC}& \text{AGC}& \text{AAC}& \text{AAU}\\ \text{GGG}& \text{AGG}& \text{AAG}& \text{AAA}\end{array}\right)`$
$`(\frac{3}{2},\frac{1}{2})^1\left(\begin{array}{cccc}\text{CCG}& \text{UCG}& \text{UUG}& \text{UUA}\\ \text{GCG}& \text{ACG}& \text{AUG}& \text{AUA}\end{array}\right)`$
$`(\frac{3}{2},\frac{1}{2})^2\left(\begin{array}{cccc}\text{CGC}& \text{UGC}& \text{UAC}& \text{UAU}\\ \text{CGG}& \text{UGG}& \text{UAG}& \text{UAA}\end{array}\right)`$
$`(\frac{1}{2},\frac{3}{2})^1\left(\begin{array}{cc}\text{CCU}& \text{UCU}\\ \text{GCU}& \text{ACU}\\ \text{GGU}& \text{AGU}\\ \text{GGA}& \text{AGA}\end{array}\right)(\frac{1}{2},\frac{3}{2})^2\left(\begin{array}{cc}\text{CUC}& \text{CUU}\\ \text{GUC}& \text{GUU}\\ \text{GAC}& \text{GAU}\\ \text{GAG}& \text{GAA}\end{array}\right)`$
$`(\frac{1}{2},\frac{1}{2})^1\left(\begin{array}{cc}\text{CCA}& \text{UCA}\\ \text{GCA}& \text{ACA}\end{array}\right)(\frac{1}{2},\frac{1}{2})^2\left(\begin{array}{cc}\text{CGU}& \text{UGU}\\ \text{CGA}& \text{UGA}\end{array}\right)`$
$`(\frac{1}{2},\frac{1}{2})^3\left(\begin{array}{cc}\text{CUG}& \text{CUA}\\ \text{GUG}& \text{GUA}\end{array}\right)(\frac{1}{2},\frac{1}{2})^4\left(\begin{array}{cc}\text{CAC}& \text{CAU}\\ \text{CAG}& \text{CAA}\end{array}\right)`$
The correspondence with the amino-acids is given in Table 10 (for the eukariotic code).
Let us close this section by drawing the readerโs attention to Fig. 1 where is specified for each codon its position in the appropriate representation. The diagram of states for each representation is supposed to lie in a separate parallel plane. Thick lines connect codons associated to the same amino-acid. One remarks that each segment relates a couple of codons belonging to the same representation or to two different representations. This last case occurs for quadruplets or sextets of codons associated to the same amino-acid.
## 3 The Reading (or Ribosome) operator $``$
### 3.1 General structure of the reading operator
As expected from formula (11), our model does not gather codons associated to one particular amino-acid in the same irreducible multiplet. However, it is possible to construct an operator $``$ out of the algebra $`๐ฐ_{q0}(sl(2)sl(2))`$, acting on the codons, that will describe the various genetic codes in the following way:
*Two codons have the same eigenvalue under $``$ if and only if they are associated to the same amino-acid. This operator $``$ will be called the reading operator.*
It is a remarkable fact that the various genetic codes share the same basic structure. As we mentioned above, the dinucleotides can be split into โstrongโ dinucleotides CC, GC, UC, AC, CU, GU, CG and GG that lead to quartets and โweakโ ones UU, AU, UG, AG, CA, GA, UA, AA that lead to doublets. Let us construct a prototype of the reading operator that reproduces this structure.
The first part of the reading operator $``$ is responsible for the structure in quadruplets given essentially by the dinucleotide content. It is given by (the $`c_i`$ are arbitrary coefficients)
$$\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}.$$
(18)
The operators $`J_{\alpha ,3}`$ ($`\alpha =H,V`$) are the third components of the total spin generators of the algebra $`๐ฐ_{q0}(sl(2)sl(2))`$. The operator $`C_\alpha `$ is a Casimir operator of $`๐ฐ_{q0}(sl(2)_\alpha )`$ in the crystal basis. It commutes with $`J_{\alpha \pm }`$ and $`J_{\alpha ,3}`$ and its eigenvalues on any vector basis of an irreducible representation of highest weight $`J`$ is $`J(J+1)`$, that is the same as the undeformed standard second degree Casimir operator of $`sl(2)`$. Its explicit expression is
$$C_\alpha =(J_{\alpha ,3})^2+\frac{1}{2}\underset{n_+}{}\underset{k=0}{\overset{n}{}}(J_\alpha )^{nk}(J_{\alpha +})^n(J_\alpha )^k.$$
(19)
Note that for $`sl(2)_{q0}`$ the Casimir operator is an infinite series of powers of $`J_{\alpha \pm }`$. However in any finite irreducible representation only a finite number of terms gives a non-vanishing contribution.
$`๐ซ_H`$ and $`๐ซ_V`$ are projectors given by the following expressions:
$$๐ซ_H=J_{H+}^dJ_H^d\text{and}๐ซ_V=J_{V+}^dJ_V^d.$$
(20)
The second part of $``$ gives rise to the splitting of the quadruplets into doublets. It reads
$$2๐ซ_Dc_3J_{V,3}$$
(21)
where the projector $`๐ซ_D`$ is given by
$`๐ซ_D`$ $`=`$ $`(1J_{V+}^dJ_V^d)(J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)+(1J_{H+}^dJ_H^d)(1J_{V+}^dJ_V^d)`$ (22)
$`+(1J_{H+}^dJ_H^d)(J_{V+}^dJ_V^d)(J_H^dJ_{H+}^d).`$
The third part of $``$ allows to reproduce the sextets viewed as quartets plus doublets. It is
$$2๐ซ_Sc_4J_{V,3}$$
(23)
where the projector $`๐ซ_S`$ is given by
$$๐ซ_S=(J_H^dJ_{H+}^d)[(J_{H+}^dJ_H^d)(1J_{V+}^dJ_V^d)+(J_{V+}^dJ_V^d)(J_V^dJ_{V+}^d)(1J_{H+}^dJ_H^d)].$$
(24)
At this point, one obtains the eigenvalues of the reading operator $``$ for the 64 codons, where Y = C,U (pyrimidines), R = G,A (purines) and N = C,U,G,A:
$$\begin{array}{ccc}\text{CCN}=c_1c_2\hfill & & \text{GCN}=c_1+3c_2\hfill \\ \text{UCN}=3c_1c_2\hfill & & \text{ACN}=3c_1+3c_2\hfill \\ \text{CUN}=c_1c_2\hfill & & \text{GUN}=c_1+3c_2\hfill \\ \text{CGN}=c_1+c_2\hfill & & \text{GGN}=c_1+5c_2\hfill \\ \text{UUY}=5c_1c_23c_3\hfill & & \text{UUR}=5c_1c_2c_3\hfill \\ \text{AUY}=5c_1+3c_2c_3c_4\hfill & & \text{AUR}=5c_1+3c_2+c_3+c_4\hfill \\ \text{UGY}=3c_1+c_2c_3c_4\hfill & & \text{UGR}=3c_1+c_2+c_3+c_4\hfill \\ \text{AGY}=3c_1+5c_2+c_3+c_4\hfill & & \text{AGR}=3c_1+5c_2+3c_3+3c_4\hfill \\ \text{CAY}=c_1+c_2c_3\hfill & & \text{CAR}=c_1+c_2+c_3\hfill \\ \text{GAY}=c_1+5c_2+c_3\hfill & & \text{GAR}=c_1+5c_2+3c_3\hfill \\ \text{UAY}=5c_1+c_2c_3\hfill & & \text{UAR}=5c_1+c_2+c_3\hfill \\ \text{AAY}=5c_1+5c_2+c_3\hfill & & \text{AAR}=5c_1+5c_2+3c_3\hfill \end{array}$$
(25)
The coefficients $`c_3`$ and $`c_4`$ are fixed as follows. The coefficient $`c_3`$ is set to the value $`c_3=4c_1`$ by requiring that the quartet CUN and the doublet UUR, associated to the amino-acid Leu, lead to the same $``$-eigenvalue. It remains to reproduce the Ser sextet. This is achieved by taking for the coefficient $`c_4`$ the value $`c_4=4c_16c_2`$, such that the final eigenvalues for the codons are the following:
$$\begin{array}{ccccccc}\text{CCN}=c_1c_2\hfill & & \text{GCN}=c_1+3c_2\hfill & & \text{UCN}=3c_1c_2\hfill & & \text{ACN}=3c_1+3c_2\hfill \\ \text{CUN}=c_1c_2\hfill & & \text{GUN}=c_1+3c_2\hfill & & \text{CGN}=c_1+c_2\hfill & & \text{GGN}=c_1+5c_2\hfill \\ \text{UUY}=7c_1c_2\hfill & & \text{UUR}=c_1c_2\hfill & & \text{AUY}=5c_1+9c_2\hfill & & \text{AUR}=5c_13c_2\hfill \\ \text{UGY}=3c_1+7c_2\hfill & & \text{UGR}=3c_15c_2\hfill & & \text{AGY}=3c_1c_2\hfill & & \text{AGR}=3c_113c_2\hfill \\ \text{CAY}=3c_1+c_2\hfill & & \text{CAR}=5c_1+c_2\hfill & & \text{GAY}=5c_1+5c_2\hfill & & \text{GAR}=13c_1+5c_2\hfill \\ \text{UAY}=c_1+c_2\hfill & & \text{UAR}=9c_1+c_2\hfill & & \text{AAY}=9c_1+5c_2\hfill & & \text{AAR}=17c_1+5c_2\hfill \end{array}$$
(26)
The prototype of the reading operator $``$ takes finally the form:
$`=\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (27)
and the correspondence codons/amino-acids is given as follows:
| CCN $``$ Pro | | UCN $``$ Ser | | GCN $``$ Ala | | ACN $``$ Thr |
| --- | --- | --- | --- | --- | --- | --- |
| CUN $``$ Leu | | GUN $``$ Val | | CGN $``$ Arg | | GGN $``$ Gly |
| UUY $``$ Phe | | AUY $``$ Ile | | UGY $``$ Cys | | AGY $``$ Ser |
| UUR $``$ Leu | | AUR $``$ Met | | UGR $``$ Trp | | AGR $``$ unassigned (X) |
| CAY $``$ His | | UAY $``$ Tyr | | GAY $``$ Gln | | AAY $``$ Asn |
| CAR $``$ Gln | | UAR $``$ Ter | | GAR $``$ Glu | | AAR $``$ Lys |
(28)
### 3.2 The various genetic codes
In this section, we will determine the reading operators for the following genetic codes:
โ the Eukariotic Code (EC),
โ the Vertebral Mitochondrial Code (VMC),
โ the Yeast Mitochondrial Code (YMC),
โ the Invertebrate Mitochondrial Code (IMC),
โ the Protozoan Mitochondrial and Mycoplasma Code (PMC),
โ the Echinoderm Mitochondrial Code (EMC),
โ the Ascidian Mitochondrial Code (AMC),
โ the Flatworm Mitochondrial Code (FMC),
โ the Ciliate Nuclear Code (CNC),
โ the Blepharisma Nuclear Code (BNC),
โ the Euplotid Nuclear Code (ENC),
โ the Alternative Yeast Nuclear Code (alt. YNC),
Let us emphasize that each of these codes is very close to the assignment (28). The main differences between the biological codes and the prototype code (28) are the following:
* assignment of the doublet AGR either to Arg (codes EC, YMC, PMC, CNC, BNC, ENC, aYNC), Ser (codes IMC, EMC, FMC), Gly (code AMC) or the stop signal Ter (code VMC).
Such an assignment is done by the following term in the reading operator:
$$c_5๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)$$
(29)
The operators $`J_{\alpha ,3}^{(3)}`$ are the third components corresponding to the third nucleotide of a codon. Of course, these last two operators can be replaced by $`J_{\alpha ,3}^{(3)}=J_{\alpha ,3}J_{\alpha ,3}^d`$.
The projector $`๐ซ_{AG}`$ is given by
$$๐ซ_{AG}=(J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)(1J_{V+}^dJ_V^d)(J_V^dJ_{V+}^d)$$
(30)
and the coefficient $`c_5`$ by
$$\begin{array}{cc}\text{for }\text{Arg}\hfill & c_5=4c_1+14c_2\hfill \\ \text{for }\text{Ser}\hfill & c_5=12c_2\hfill \\ \text{for }\text{Gly}\hfill & c_5=4c_1+18c_2\hfill \\ \text{for }\text{Ter}\hfill & c_5=6c_1+14c_2\hfill \end{array}$$
(31)
* splitting of some doublets into singlets (one element of the singlet combining to another doublet to form a triplet):
Met $``$ Met \+ Ile for the EC, PMC, EMC, FMC, CNC, BNC, ENC, aYNC codes;
Lys $``$ Lys \+ Asn for the FMC and EMC codes;
Trp $``$ Trp \+ Ter for the EC, CNC, BNC, aYNC codes;
Trp $``$ Trp \+ Cys for the ENC code;
Ter $``$ Tyr \+ Ter for the FMC code;
Such an assignment is done through the following term in the reading operator:
$$c_6๐ซ_{XY}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)$$
(32)
where we use the projector $`๐ซ_{AU}`$ for the splitting of the Met doublet, $`๐ซ_{AA}`$ for the Lys doublet, $`๐ซ_{UG}`$ for the Trp doublet, and $`๐ซ_{UA}`$ for the Ter doublet. These projectors are given by
$`๐ซ_{AU}=(1J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)(J_{V+}^dJ_V^d)(J_V^dJ_{V+}^d)`$ (33)
$`๐ซ_{AA}=(1J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)(1J_{V+}^dJ_V^d)(J_V^dJ_{V+}^d)`$ (34)
$`๐ซ_{UG}=(J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)(1J_{V+}^dJ_V^d)(1J_V^dJ_{V+}^d)`$ (35)
$`๐ซ_{UA}=(1J_{H+}^dJ_H^d)(J_H^dJ_{H+}^d)(1J_{V+}^dJ_V^d)(1J_V^dJ_{V+}^d)`$ (36)
The coefficient $`c_6`$ takes the following values:
$$\begin{array}{cc}\text{for }\text{Met}\text{ }\text{ }\text{Met}\text{ + }\text{Ile}\hfill & c_6=12c_2\hfill \\ \text{for }\text{Lys}\text{ }\text{ }\text{Lys}\text{ + }\text{Asn}\hfill & c_6=8c_1\hfill \\ \text{for }\text{Trp}\text{ }\text{ }\text{Trp}\text{ + }\text{Cys}\hfill & c_6=12c_2\hfill \\ \text{for }\text{Trp}\text{ }\text{ }\text{Trp}\text{ + }\text{Ter}\hfill & c_6=6c_1+6c_2\hfill \\ \text{for }\text{Ter}\text{ }\text{ }\text{Ter}\text{ + }\text{Tyr}\hfill & c_6=8c_1\hfill \end{array}$$
(37)
* in the case of the CNC and BNC codes, the Ter doublet is changed in Gln as follows:
Ter $``$ Gln for the CNC code by the term
$$4c_1๐ซ_{UA}\left(\frac{1}{2}J_{V,3}^{(3)}\right)$$
(38)
Ter $``$ Ter \+ Gln for the BNC code by the term
$$4c_1๐ซ_{UA}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}+J_{H,3}^{(3)}\right)$$
(39)
* in the case of the alternative YNC code, the last quartet Leu is split into a triplet Leu coded by (CUC,CUU,CUA) and a doublet Ser coded by (CUG). The corresponding term in the reading operator is
$$2c_1๐ซ_{CU}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}+J_{H,3}^{(3)}\right)$$
(40)
where the projector $`๐ซ_{CU}`$ is given by
$$๐ซ_{CU}=(1J_{H+}^dJ_H^d)(1J_H^dJ_{H+}^d)(J_{V+}^dJ_V^d)(1J_V^dJ_{V+}^d)$$
(41)
* in the case of the Yeast Mitochondrial Code, the quartet CUN codes the amino-acid Thr rather than Leu. This change is achieved by multiplying the quartets term (18) by $`(1+2๐ซ_{CU})`$ for the horizontal part and by $`(14๐ซ_{CU})`$ for the vertical part.
#### 3.2.1 The Eukariotic Code (EC)
The Eukariotic Code is the most important one and is often referred to as the universal code. The differences between the Eukariotic Code and the prototype code are the following:
| | prototype code | EC | | | | prototype code | EC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
| UGG | Trp | Trp | | | UGA | Trp | Ter |
Hence from (29), (31), (32) and (37), the reading operator for the Eukariotic Code is
$`_{EC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (42)
$`+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
$`+\left[12c_2๐ซ_{AU}+(6c_1+6c_2)๐ซ_{UG}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.2 The Vertebral Mitochondrial Code (VMC)
The Vertebral Mitochondrial Code is used in the mitochondriae of vertebrata. The differences between the Vertebral Mitochondrial Code and the prototype code are the following:
| | prototype code | VMC | | | | prototype code | VMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AGG | X | Ter | | | AGA | X | Ter |
Hence from (29) and (31), the reading operator for the Vertebral Mitochondrial Code is
$`_{VMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (43)
$`+(6c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
#### 3.2.3 The Yeast Mitochondrial Code (YMC)
The Yeast Mitochondrial Code is used in the mitochondriae of yeast such as Saccharomyces, Candida, etc. The differences between the Yeast Mitochondrial Code and the prototype code are the following:
| | prototype code | YMC | | | | prototype code | YMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| CUC | Leu | Thr | | | CUU | Leu | Thr |
| CUG | Leu | Thr | | | CUA | Leu | Thr |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29) and (31), the reading operator for the Yeast Mitochondrial Code is
$`_{YMC}`$ $`=`$ $`(\frac{4}{3}c_1C_H4c_1๐ซ_HJ_{H,3})(1+2๐ซ_{CU})+(\frac{4}{3}c_2C_V4c_2๐ซ_VJ_{V,3})(14๐ซ_{CU})`$ (44)
$`+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
#### 3.2.4 The Invertebrate Mitochondrial Code (IMC)
The Invertebrate Mitochondrial Code is used in the mitochondriae of some arthopoda, mollusca, nematoda and insecta. The differences between the Invertebrate Mitochondrial Code and the prototype code are the following:
| | prototype code | IMC | | | | prototype code | IMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AGG | X | Ser | | | AGA | X | Ser |
Hence from (29) and (31), the reading operator for the Invertebrate Mitochondrial Code is
$`_{IMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (45)
$`+12c_2๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
#### 3.2.5 The Protozoan Mitochondrial and Mycoplasma Code (PMC)
The Protozoan Mitochondrial and Mycoplasma Code is used in the mitochondriae of some protozoa (leishmania, paramecia, trypanosoma, etc.) and for many fungi. The differences between the Protozoan Mitochondrial and Mycoplasma Code and the prototype code are the following:
| | prototype code | PMC | | | | prototype code | PMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29), (31), (32) and (37), the reading operator for the Protozoan Mitochondrial Code is
$`_{PMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (46)
$`+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)+12c_2๐ซ_{AU}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.6 The Echinoderm Mitochondrial Code (EMC)
The Echinoderm Mitochondrial Code is used in the mitochondriae of some asterozoa and echinozoa. The differences between the Echinoderm Mitochondrial Code and the prototype code are the following:
| | prototype code | EMC | | | | prototype code | EMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Ser | | | AGA | X | Ser |
| AAG | Lys | Lys | | | AAA | Lys | Asn |
Hence from (29), (31), (32) and (37), the reading operator for the Echinoderm Mitochondrial Code is
$`_{EMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (47)
$`+12c_2๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)+\left[12c_2๐ซ_{AU}8c_1๐ซ_{AA}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.7 The Ascidian Mitochondrial Code (AMC)
The Ascidian Mitochondrial Code is used in the mitochondriae of some ascidiacea. The differences between the Ascidian Mitochondrial Code and the prototype code are the following:
| | prototype code | AMC | | | | prototype code | AMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| AGG | X | Gly | | | AGA | X | Gly |
Hence from (29) and (31), the reading operator for the Ascidian Mitochondrial Code is
$`_{AMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (48)
$`+(4c_1+18c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
#### 3.2.8 The Flatworm Mitochondrial Code (FMC)
The Flatworm Mitochondrial Code is used in the mitochondriae of the flatworms. The differences between the Flatworm Mitochondrial Code and the prototype code are the following:
| | prototype code | FMC | | | | prototype code | FMC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| UAG | Ter | Ter | | | UAA | Ter | Tyr |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Ser | | | AGA | X | Ser |
| AAG | Lys | Lys | | | AAA | Lys | Asn |
Hence from (29), (31), (32) and (37), the reading operator for the Flatworm Mitochondrial Code is
$`_{FMC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$
$`+12c_2๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)+\left[12c_2๐ซ_{AU}8c_1๐ซ_{AA}8c_1๐ซ_{UA}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.9 The Ciliate Nuclear Code (CNC)
The Ciliate Nuclear Code is used in the nuclei of some ciliata, dasyclasaceae and diplomonadida. The differences between the Ciliate Nuclear Code and the prototype code are the following:
| | prototype code | CNC | | | | prototype code | CNC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| UGG | Trp | Trp | | | UGA | Trp | Ter |
| UAG | Ter | Gln | | | UAA | Ter | Gln |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29), (31), (32), (37) and (38), the reading operator for the Ciliate Nuclear Code is
$`_{CNC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (50)
$`+\left[(4c_1+14c_2)๐ซ_{AG}4c_1๐ซ_{UA}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)`$
$`+\left[12c_2๐ซ_{AU}+(6c_1+6c_2)๐ซ_{UG}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.10 The Blepharisma Nuclear Code (BNC)
The Blepharisma Nuclear Code is used in the nuclei of the blepharisma (ciliata) (note that this code is very close to the CNC which is used for the ciliata). The differences between the Blepharisma Nuclear Code and the prototype code are the following:
| | prototype code | BNC | | | | prototype code | BNC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| UGG | Trp | Trp | | | UGA | Trp | Ter |
| UAG | Ter | Gln | | | UAA | Ter | Ter |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29), (31), (32), (37) and (39), the reading operator for the Blepharisma Nuclear Code is
$`_{BNC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (51)
$`+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)4c_1๐ซ_{UA}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}+J_{H,3}^{(3)}\right)`$
$`+\left[12c_2๐ซ_{AU}+(6c_1+6c_2)๐ซ_{UG}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.11 The Euplotid Nuclear Code (ENC)
The Euplotid Nuclear Code is used in the nuclei of the euplotidae (ciliata). The differences between the Euplotid Nuclear Code and the prototype code are the following:
| | prototype code | ENC | | | | prototype code | ENC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| UGG | Trp | Trp | | | UGA | Trp | Cys |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29), (31), (32) and (37), the reading operator for the Euplotid Nuclear Code is
$`_{ENC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (52)
$`+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)+12c_2(๐ซ_{AU}+๐ซ_{UG})\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
#### 3.2.12 The alternative Yeast Nuclear Code (alt. YNC)
The alternative Yeast Nuclear Code is used in the nuclei of some yeast (essentially many candidae). The differences between the alternative Yeast Nuclear Code and the prototype code are the following:
| | prototype code | alt. YNC | | | | prototype code | alt. YNC |
| --- | --- | --- | --- | --- | --- | --- | --- |
| CUG | Leu | Ser | | | CUA | Leu | Leu |
| UGG | Trp | Trp | | | UGA | Trp | Ter |
| AUG | Met | Met | | | AUA | Met | Ile |
| AGG | X | Arg | | | AGA | X | Arg |
Hence from (29), (31), (32), (37) and (40), the reading operator for the alternative Yeast Nuclear Code is
$`_{aYNC}`$ $`=`$ $`\frac{4}{3}c_1C_H+\frac{4}{3}c_2C_V4c_1๐ซ_HJ_{H,3}4c_2๐ซ_VJ_{V,3}+(8c_1๐ซ_D+(8c_1+12c_2)๐ซ_S)J_{V,3}`$ (53)
$`+(4c_1+14c_2)๐ซ_{AG}\left(\frac{1}{2}J_{V,3}^{(3)}\right)+2c_1๐ซ_{CU}\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}+J_{H,3}^{(3)}\right)`$
$`+\left[(6c_1+6c_2)๐ซ_{UG}+12c_2๐ซ_{AU}\right]\left(\frac{1}{2}J_{V,3}^{(3)}\right)\left(\frac{1}{2}J_{H,3}^{(3)}\right)`$
### 3.3 Reading values for the amino-acids
We have therefore constructed reading operators for the genetic codes specified above, starting from a prototype code that emphasizes the quartet/doublet structure of the different codes. The different reading operators are such that they give the same value for a given amino-acid, whatever the code under consideration. Finally, we get the following eigenvalues of the reading operators for the amino-acids (after a rescaling, setting $`cc_1/c_2`$):
$$\begin{array}{cccccc}\text{a.a.}& \hfill \text{value of }& \text{a.a.}& \hfill \text{value of }& \text{a.a.}& \hfill \text{value of }\\ & & & & & \\ .\text{Ala}& \hfill c+3& .\text{Gly}& \hfill c+5& .\text{Pro}& \hfill c1\\ .\text{Arg}& \hfill c+1& .\text{His}& \hfill 3c+1& .\text{Ser}& \hfill 3c1\\ .\text{Asn}& \hfill 9c+5& .\text{Ile}& \hfill 5c+9& .\text{Thr}& \hfill 3c+3\\ .\text{Asp}& \hfill 5c+5& .\text{Leu}& \hfill c1& .\text{Trp}& \hfill 3c5\\ .\text{Cys}& \hfill 3c+7& .\text{Lys}& \hfill 17c+5& .\text{Tyr}& \hfill c+1\\ .\text{Gln}& \hfill 5c+1& .\text{Met}& \hfill 5c3& .\text{Val}& \hfill c+3\\ .\text{Glu}& \hfill 13c+5& .\text{Phe}& \hfill 7c1& .\text{Ter}& \hfill 9c+1\end{array}$$
(54)
Remark that the reading operators $`(c)`$ can be used for any real value of $`c`$, except those conferring the same eigenvalue to codons relative to two different amino-acids. These forbidden values are the following: $`7`$, $`5`$, $`4`$, $`3`$, $`\frac{5}{2}`$, $`\frac{7}{3}`$, $`2`$, $`\frac{5}{3}`$, $`\frac{3}{2}`$, $`\frac{4}{3}`$, $`1`$, $`\frac{5}{6}`$, $`\frac{4}{5}`$, $`\frac{3}{4}`$, $`\frac{5}{7}`$, $`\frac{2}{3}`$, $`\frac{3}{5}`$, $`\frac{1}{2}`$, $`\frac{3}{7}`$, $`\frac{2}{5}`$, $`\frac{3}{8}`$, $`\frac{1}{3}`$, $`\frac{3}{10}`$, $`\frac{2}{7}`$, $`\frac{1}{4}`$, $`\frac{2}{9}`$, $`\frac{1}{5}`$, $`\frac{1}{6}`$, $`\frac{1}{7}`$, $`\frac{1}{8}`$, $`\frac{1}{9}`$, $`0`$, $`\frac{1}{7}`$, $`\frac{1}{6}`$, $`\frac{1}{5}`$, $`\frac{1}{4}`$, $`\frac{1}{3}`$, $`\frac{2}{5}`$, $`\frac{1}{2}`$, $`\frac{2}{3}`$, $`1`$, $`\frac{4}{3}`$, $`\frac{3}{2}`$, $`2`$, $`\frac{5}{2}`$, $`3`$, $`4`$, $`5`$.
At this point, let us emphasize the specific properties of our model. To each nucleotide are assigned specific quantum numbers characterizing its purine/pyrimidine origin and involving the complementary rule. Then ordered sequences of bases can be constructed and characterized in this framework. Ordered sequences of three bases have been just above examined and the correspondence codon/amino-acid represented by the reading operator $``$. Finally let us remark that the coefficients $`c_i`$, which above have been taken as constants, can more generally be considered as functions of some external variables (biological, physical and chemical environment, time, etc.). In this way it is possible to explain the observed discrepancy in the correspondence codons/amino-acid in biological species under stress conditions (in vitro). In this scheme the evolution process of genetic code can also be discussed. However, we believe that a better understanding of the reasons of the evolution, i.e. which kind of optimization process takes place, has still to be acquired.
## 4 Physical properties of the dinucleotides
The model we have at hand, with nucleotides characterized by quantum numbers, is well adapted to elaborate formulae expressing biophysical properties. A particularly interesting quantity is the free energy released by base pairing in double stranded RNA. The data are not provided for a doublet of nucleotides, with one item in each strand, but for a pair of nucleotides, for ex. CG, lying on one strand and coupled with another pair, i.e. GC on the second strand ; note also that the direction on a strand being perfectly defined, the release of energy for the doublet sequence CG on the first strand running from $`5^{}`$ to $`3^{}`$ related to the doublet GC on the complementary strand running from $`3^{}`$ to $`5^{}`$, will be different to the one related to the doublet GC, itself associated to CG. It appears clear that such quantities involve pairs of nucleotides, and that naturally ordered crystal bases obtained from tensor product of two representations are adapted for such a calculation.
We will also consider two other quantities involving again pairs of nucleotides, namely the relative hydrophilicity $`R_f`$ and hydrophobicity $`R_x`$ of dinucleosides.
Before presenting our results, let us mention that fits for the same biophysical properties can be found in a recent preprint where polynomials in 4 or 6 coordinates in the 64 codon space are constructed. In their approach, the authors associate two coordinates $`(d,m)`$ to each nucleotide of any codon, as follows: $`A=(1,0)`$, $`C=(0,1)`$, $`G=(0,1)`$, $`U=(1,0)`$, labelling in this way each codon with 6 numbers. The above labelling of the nucleotides is related to our labels Eq. (LABEL:eq:gc1) in the following way:
$$\begin{array}{ccc}& d& m\\ & & \\ C& J_{V,3}J_{H,3}& (J_{V,3}+J_{H,3})\\ U& J_{V,3}J_{H,3}& (J_{V,3}+J_{H,3})\\ G& J_{V,3}+J_{H,3}& J_{V,3}J_{H,3}\\ A& J_{V,3}+J_{H,3}& J_{V,3}J_{H,3}\end{array}$$
Therefore the labels $`(d,m)`$ just correspond up to a sign for the pyrimidine (resp. purine) to the antidiagonal and diagonal (resp. diagonal and antidiagonal) $`๐ฐ_{q0}(sl(2))`$.
In the following we compare our results with those of .
#### Free energy
In we have fitted the experimental data with a four-parameter operator. Here we fit the more recent data with a two-parameter operator obtained from the one used in by setting two parameters to zero:
$$\mathrm{\Delta }G_{37}^0=\alpha _0+\alpha _1(C_H+C_V)J_{3H}^d$$
(55)
Using a least-squares fit, one finds for the coefficients $`\alpha _i`$:
$$\alpha _0=2.14,\alpha _1=0.295$$
(56)
The standard deviation of the two-parameter fit (56) is found to be equal to $`0.149`$, which is to be compared to the standard deviation $`0.16`$ of the four-parameter fit of ref. . The experimental and fitted values of the free energies $`\mathrm{\Delta }G_{37}^0`$ of the dinucleotides are displayed in Table 1.
#### Hydrophilicity
We fit the values of the relative hydrophilicity $`R_f`$ of the 16 dinucleoside monophosphates with the following four-parameter operator:
$$R_f=\alpha _0+\alpha _1C_V+\alpha _2J_{3V}^d+\alpha _3\underset{i=1,2}{}(J_{3H}^i+J_{3V}^i)(J_{3H}^i+J_{3V}^i1)$$
(57)
(the last term in $`\alpha _3`$ is equal to 4 for AA, to 2 for CA, GA, UA and zero for the other dinucleotides).
Using a least-squares fit, one finds for the coefficients $`\alpha _i`$:
$$\alpha _0=0.135,\alpha _1=0.036,\alpha _2=0.147,\alpha _3=0.016$$
(58)
The standard deviation of the four-parameter fit (58) is found to be equal to $`0.027`$, which is to be compared to the standard deviation $`0.033`$ of the six-parameter fit of ref. . The experimental and fitted values of the hydrophilicity $`R_f`$ of the dinucleosides are displayed in Table 2.
#### Hydrophobicity
We fit the values of the relative hydrophobicity $`R_x`$ of the 16 dinucleoside monophosphates as reported in with the following four-parameter operator:
$$R_x=\alpha _0+\alpha _1J_{3V}^d+\alpha _2J_{3H}^d+\alpha _3[(J_{3H}^1+J_{3V}^1)^2+(J_{3H}^2+J_{3V}^2)^2]$$
(59)
(the last term in $`\alpha _3`$ is equal to 2 for AA, AC, CA and CC, to 2 for AU, AG, UA, UC, GC, GA, CU and CG and zero for UU, UG, GU, GG).
Using a least-squares fit, one finds for the coefficients $`\alpha _i`$:
$$\alpha _0=0.294,\alpha _1=0.240,\alpha _2=0.105,\alpha _3=0.136$$
(60)
Using a least-squares fit without the dinucleoside AA, one finds new coefficients $`\alpha _i`$, which lead to better values of $`R_x`$ for the remaining dinucleosides:
$$\alpha _0=0.309,\alpha _1=0.203,\alpha _2=0.068,\alpha _3=0.099$$
(61)
The standard deviation of the four-parameter fit (60) is equal to $`0.049`$, which is the same of the four-parameter fit of ref. . Using the fit (61), the standard deviation becomes $`0.074`$ (including the value for AA) or $`0.024`$ (excluding the value for AA). For this last case, the standard deviation of ref. is still equal to $`0.031`$. The experimental and fitted values (second fit) of the relative hydrophobicity $`R_x`$ of the dinucleosides are displayed in Table 3.
## 5 Universal behaviour of ratios of codon usage frequency
In the following the labels $`X,J,Z,K`$ represent any of the 4 bases $`C,U,G,A`$. Let $`XJZ`$ be a codon in a given multiplet, say $`m_i`$, encoding an a.a., say $`A_i`$. We define the probability of usage of the codon $`XJZ`$ as the ratio between the frequency of usage $`n_Z`$ of the codon $`XJZ`$ in the biosynthesis of $`A_i`$ and the total number $`n`$ of synthesized $`A_i`$, i.e. as the relative codon frequency, in the limit of *very large $`n`$*.
It is natural to assume that the usage frequency of a codon in a multiplet is connected to its probability of usage $`P(XJZA_i)`$. We define the *branching ratio* $`B_{ZK}`$ as
$$B_{ZK}=\frac{P(XJZA_i)}{P(XJKA_i)}$$
(62)
where $`XJK`$ is another codon belonging to the same multiplet $`m_i`$. It is reasonable to argue that in the limit of very large number of codons, for a fixed biological species and amino-acid, the branching ratio depends essentially on the properties of the codon. In our model this means that in this limit $`B_{ZK}`$ is a function, depending on the type of the multiplet, on the *quantum numbers* of the codons $`XJZ`$ and $`XJK`$, i.e. on the labels $`J_\alpha ,J_{\alpha ,3}`$, where $`\alpha =H`$ or $`V`$, and on an other set of quantum labels leaving out the degeneracy on $`J_\alpha `$; in Table 4 different irreducible representations with the same values of $`J_\alpha `$ are distinguished by an upper label.
We have put in evidence a correlation in the codon usage frequency for the quartets and the quartet subpart of the sextets, i.e. the codons in a sextet differing only for the third codon, for the vertebrates in and for biological species belonging to the vertebrates, invertebrates, plants and fungi in , and we have shown that these correlations fit well in our model with the assumed dependence on $`B_{ZK}`$. Here we remark that for thirteen biological species belonging to the vertebrate class, with a statistics of codons larger than 95,000 (see Table 5), the ratio of
$$\frac{B_{AG}}{B_{UC}}=\frac{B_{AU}}{B_{GC}}=\frac{P(XJAA_i)}{P(XJGA_i)}\frac{P(XJCA_i)}{P(XJUA_i)}$$
(63)
for quartets and the quartet subpart of the sextets has a behaviour independent of the specific biological species. Moreover, for the same amino-acids for which we have remarked correlations, the values of the ratio $`B_{AG}/B_{UC}`$ are almost the same (see Table 8). We show that these behaviour and correlations find a nice explanation in our model. In Tables 6 and 7, we report respectively the values of the branching ratios $`B_{AG}`$ and $`B_{UC}`$ as computed from the database (release of February 2000) and in Table 8 the ratio of these quantities. The average values $`B_{AG}/B_{UC}`$, the standard deviations $`\sigma `$ and the ratios $`\sigma /B_{AG}/B_{UC}`$ are displayed in the following table:
$$\begin{array}{ccccccccc}& & & & & & & & \\ & \hfill \mathrm{Pro}& \hfill \mathrm{Ala}& \hfill \mathrm{Thr}& \hfill \mathrm{Ser}& \hfill \mathrm{Gly}& \hfill \mathrm{Val}& \hfill \mathrm{Leu}& \hfill \mathrm{Arg}\\ & & & & & & & & \\ B_{AG}/B_{UC}\hfill & \hfill 2.50& \hfill 2.84& \hfill 3.30& \hfill 2.67& \hfill 2.21& \hfill 0.33& \hfill 0.26& \hfill 1.32\\ \sigma \hfill & \hfill 0.46& \hfill 0.53& \hfill 0.56& \hfill 0.35& \hfill 0.30& \hfill 0.04& \hfill 0.03& \hfill 0.14\\ \sigma /B_{AG}/B_{UC}\hfill & \hfill 0.19& \hfill 0.19& \hfill 0.17& \hfill 0.13& \hfill 0.14& \hfill 0.13& \hfill 0.10& \hfill 0.11\end{array}$$
The above behaviour can be easily understood considering a dependence on $`B_{ZK}`$ not only on the irreducible representations to which the codons $`XJZ`$ and $`XJK`$ appearing in the numerator and the denominator belong, but also on the specific states denoting these codons, and refining the factorized form of as
$$B_{ZK}=F_{ZK}(IR(XJZ);IR(XJK))\frac{G_H(b.s.;J_{H,3}(XJZ))G_V(b.s.;J_{V,3}(XJZ))}{G_H(b.s.;J_{H,3}(XJK))G_V(b.s.;J_{H,3}(XJK))}$$
(64)
where we have denoted by $`b.s.`$ the biological species, by $`IR(XJZ)`$ and $`J_{\alpha ,3}(XJZ)`$ the irreducible representation to which the codon $`XJZ`$ belong (see Table 4), and the value of the third component of the $`\alpha `$-spin of the state $`XJZ`$. Note that we have still neglected the dependence on the type of the biosynthetized amino-acid. *The ratio $`B_{AG}/B_{UC}`$ using Eq. (64), is no more depending on the biological species but only on the value of the irreducible representations of the codons*. Moreover, for Pro, Ala, Thr, Ser, (resp. Val and Leu), the irreducible representations appearing in the $`F`$ functions are the same as can be seen from Table 9, so we expect the same value for the ratio, which is indeed the case (see above Table), the value of $`B_{AG}/B_{UC}`$ for the first four amino-acids (resp. for the last two amino-acids) lying in the range $`2.90\pm 15\%`$ (resp. $`0.30\pm 15\%`$). These values should be compared with the value 1.32 for Arg and 2.21 for Gly.
Let us end this section by the following remark. From the above table, one might be tempted to consider the value of the ratio $`B_{AG}/B_{UC}`$ for Gly of the same order of magnitude as the ones for Pro, Ala, Thr, Ser. Then one distinguishes, following this ratio, three groups of codons quartets: the one associated to the five just mentioned amino-acids, another one relative to Val and Leu, and a last one with Arg. Now, let us look at the dinucleotide pairs constituting the first two nucleotides in a codon in the light of our results of sect. 2: the pairs CC, GC, AC, UC and GG relative to Pro, Ala, Thr, Ser, and Gly respectively belong to the representation (1,1) of $`๐ฐ_{q0}(sl(2)sl(2))`$; the states GU and CU relative to Val and Leu respectively belong to the representation (0,1); finally CG relative to Arg also lies in a different representation (1,0).
## 6 Mutations in the genetic code
In this section, we present a mathematical framework to describe the single-base deletions in the genetic code. In starting from the observation that the single-base deletions in DNA, which occur far more frequently that single base additions, take place in the opposite site to a purine R, (R = G, A) i.e. a pyrimidine Y (Y = C, U/T) is deleted, arguments have been presented to explain why the Stop codons have the structure they have, see Table 4. We refer to the paper for more details and for references to the biological literature on the subject and we recall here just the main ideas and conclusions of . The starting point is the observed fact that deletions occur more frequently in the following sequences: YR, TTR, YTG and TR. In ref. all these sequences have been refined as YTRV, (V = C, A, G). Starting from the structure of this dangerous sequence and using the complementarity property, an analysis shows that four codons โ TAA, TAG, TTA, CTA โ are both potential deletion site codons and reverse-complementary potential site codons. As a mutation at the end of a protein chain just implies the addition of further peptides, the authors conclude that the assignment of codons TAA and TAG as Stop codons minimizes the possible deleterious effects of deletion. Indeed the codon usage frequency of the dangerous codon CTA, as it can be seen from fig. (5) of and from fig. (2) of , is very low. An analysis of the codon usage frequency exhibits an analogous behaviour for the codon TTA.
The mechanism by which the above specified sequences are preferred in the deletion process is unclear. In the following we will present a mathematical scheme in which these properties can be settled. Let us recall that the Wigner-Eckart theorem, has been extended to the quantum algebra $`U_q(sl(n))`$, and recently in to the case of $`U_{q0}(sl(2))`$.
In ($`q0`$)-tensor operators have been introduced, called crystal tensor operators, which transform as
$$J_3(\tau _m^j)m\tau _m^jJ_\pm (\tau _m^j)\tau _{m\pm 1}^j$$
(65)
Clearly, if $`|m|>j`$ then $`\tau _m^j`$ has to be considered vanishing.
The ($`q0`$)-Wigner-Eckart theorem can be written ($`j_1j`$)
$`\tau _m^j|j_1m_1`$ $`=`$ $`(1)^{2j}{\displaystyle \underset{\alpha =0}{\overset{2j}{}}}j_1+j\alpha \tau ^jj_1|j_1+j\alpha ,m_1+m`$ (66)
$`(\delta _{m_1,j_1\alpha }+\delta _{m,j\alpha }\delta _{m_1,j_1\alpha }\delta _{m,j\alpha })`$
The ($`q0`$)-Wigner-Eckart theorem has the peculiar feature that the selection rules do not depend only on the rank of the tensor operator and on the initial state, but in a crucial way from the specific component of the tensor in consideration. The tensor product of two irreducible representations in the crystal basis is not commutative (see sect. 2), therefore one has to specify which is the first representation. In the following, as in , the crystal tensor operator has to be considered as the first one.
Let us also remark the following peculiar property of crystal basis which will be used in the following. We specify it only for the case we are interested in, but it is a completely general property.
An ordered sequence, or chain, of $`n`$ nucleotides is a state belonging to an irreducible representation of $`U_{q0}((sl(2)sl(2))`$ appearing in the $`n`$-fold product of the fundamental irreducible representation $`(1/2,1/2)`$. Moreover the same property holds for any subsequence of $`m`$ ($`m<n`$) nucleotides. We can mimick the deletion of a $`N`$ nucleotide in a generic position of a coding sequence by a local annihilation operator of the $`N`$ nucleotide. In order to take into account the observed fact that the deletion of the nucleotide depends on the nature of the neighboring nucleotides, we require the annihilation operator to behave as a defined crystal tensor operator under $`U_{q0}(sl(2))_V`$ or $`U_{q0}(sl(2))_H`$ or both. In our mathematical description we have to specify the action of the annihilation operator on a chain of nucleotides. If we assume that the annihilation of the $`N`$ nucleotide behaves e.g. as a spinor crystal operator for the $`U_{q0}(sl(2))_V`$, we have to require that the deletion of the $`N`$ nucleotide from the initial chain of $`K`$ nucleotides, described by the state $`|J_i,M_i;\mathrm{\Omega }_i`$, leading to the final chain of $`K1`$ nucleotides, described by the state $`|J_f,M_f;\mathrm{\Omega }_f`$, is compatible with the ($`q0`$)-Wigner-Eckart theorem prescription for the action of the definite crystal spinor operator between the initial state $`|J_i,M_i;\mathrm{\Omega }_i`$ and the final state $`|J_f,M_f;\mathrm{\Omega }_f`$, where we have denoted by $`\mathrm{\Omega }`$ the set of all the labels necessary to identify completely the state. As we shall see, this is far from being trivial and will put constraints on the type of nucleotides surrounding the nucleotide $`N`$. We have to specify which chain has to be considered in order to study the action of the crystal tensor operator. It seems reasonable to take into account chains formed by $`K=2`$ and 3 nucleotides starting from $`N`$ in the sense of the reading of the codon sequence. So we are defining on the chain the action of a โmatrioskaโ crystal tensor operator. We assume:
Assumption : The biological mechanism responsible for the deletion of a pyrimidine C (resp. U ) in a sequence can be schematized by a local crystal tensor operator $`\tau _{1/2}^{1/2}`$ for $`U_{q0}(sl(2)_V)`$ and $`\tau _{1/2}^{1/2}`$ (resp. $`\tau _{1/2}^{1/2}`$ ) for $`U_{q0}(sl(2)_H)`$, which transforms the state YX (resp. YXZ) into the state X (resp. XZ), X, Z being any nucleotide.
By โlocal crystal tensor operatorโ we mean an operator which, in the sequence of RNA, acts on the $`K`$-chain ($`K=2,3`$) starting with Y, deleting the pyrimidine, according to the selection rules imposed by the assumed type of the crystal tensor.
Let us point out that, differently to ref. , where the DNA sequence was analyzed, we consider the transcripted RNA sequence and the deletion in the trascription of a Y.
There are 8 possible cases (we denote the initial and final states with the notation of sect. 2 and by A (resp. F) the allowed (resp. forbidden) transition). We analyze the deletion of a C (on the left) and of an U (on the right).
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,1)(\frac{1}{2},\frac{1}{2})`$ | | |
| CC | C | FโF |
| $`{}_{.}{}^{.}(0,1)(\frac{1}{2},\frac{1}{2})`$ | | |
| CU | U | AโF |
| $`{}_{.}{}^{.}(1,0)(\frac{1}{2},\frac{1}{2})`$ | | |
| CG | G | FโA |
| $`{}_{.}{}^{.}(0,0)(\frac{1}{2},\frac{1}{2})`$ | | |
| CA | A | AโA |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,1)(\frac{1}{2},\frac{1}{2})`$ | | |
| UC | C | AโF |
| UU | U | AโF |
| $`{}_{.}{}^{.}(1,0)(\frac{1}{2},\frac{1}{2})`$ | | |
| UG | G | AโA |
| UA | A | AโA |
So for the transition for the state of dinucleotide to one nucleotide state, from the assumed nature of the crystal tensor operator, it follows that a pyrimidine can be deleted if followed by a purine. Now let us consider what happens if we consider the transition from a trinucleotide to a dinucletide state. Using the previous result we consider only the state in which a purine is in second position so we have to consider 16 cases:
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(\frac{1}{2},\frac{1}{2})_{}^{4}(1,1)`$ | | |
| CAC | AC | AโA |
| CAU | AU | AโA |
| CAA | AA | AโA |
| CAG | AG | AโA |
| $`{}_{.}{}^{.}(\frac{3}{2},\frac{1}{2})_{}^{2}(1,1)`$ | | |
| CGC | GC | FโA |
| CGG | GG | FโA |
| $`{}_{.}{}^{.}(\frac{1}{2},\frac{1}{2})_{}^{2}(0,1)`$ | | |
| CGU | GU | FโA |
| CGA | GA | FโA |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(\frac{3}{2},\frac{1}{2})_{}^{2}(1,1)`$ | | |
| UAC | AC | AโA |
| UAU | AU | AโA |
| UAA | AA | AโA |
| UAG | AG | AโA |
| $`{}_{.}{}^{.}(\frac{3}{2},\frac{1}{2})_{}^{2}(1,1)`$ | | |
| UGC | GC | AโA |
| UGG | GG | AโA |
| $`{}_{.}{}^{.}(\frac{1}{2},\frac{1}{2})_{}^{2}(0,1)`$ | | |
| UGU | GU | AโA |
| UGA | GA | AโA |
So, from the assumed nature of the crystal tensor operator, the transition from a trinucleotide to a dinucleotide state is horizontally forbidden for the deletion of a C if the second nucleotide is a G.
Let us note that we have made the simplified assuption that the transitions depend only on the values of $`J_\alpha ,J_{\alpha ,3}`$ of the initial and final state.
Moreover, both to take into account the data of and to check that the results are not very sensible to the choice of the initial state, we consider the deletion of a purine in second position in a four-nucleotide state and impose that the process may take place only if the initial and final state can be connected by a spinor crystal operator $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ for the deletion of C or $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ for the deletion of U.
As the two pyrimidines differ by their value of $`J_{H,3}`$, the constraints imposed by the tensor operator $`\tau _{\pm 1/2,H}^{1/2}`$ are weaker than those imposed by the tensor operator $`\tau _{1/2,V}^{1/2}`$.
In Appendix (in sect. 2) we have reported all the irreducible representations arising by the $`4`$-fold ($`3`$-fold) tensor product of the fundamental representation. A detailed analysis shows that only the following deletions may happen (we report all the transitions that are allowed at least once):
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(2,1)_{}^{3}(\frac{3}{2},\frac{3}{2})`$ | | |
| GCGC | GGC | FโA |
| ACGC | AGC | FโA |
| GCGG | GGG | FโA |
| ACGG | AGG | FโA |
| $`{}_{.}{}^{.}(2,0)_{}^{2}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| CCGG | CGG | FโA |
| UCGG | UGG | FโA |
| $`{}_{.}{}^{.}(1,1)_{}^{7}(\frac{1}{2},\frac{3}{2})^1`$ | | |
| GCGU | GGU | FโA |
| ACGU | AGU | FโA |
| GCGA | GGA | FโA |
| ACGA | AGA | FโA |
| $`{}_{.}{}^{.}(1,0)_{}^{4}(\frac{1}{2},\frac{1}{2})^2`$ | | |
| CCGA | CGA | FโA |
| UCGA | UGA | FโA |
| $`{}_{.}{}^{.}(1,1)_{}^{9}(\frac{1}{2},\frac{3}{2})^2`$ | | |
| GCAC | GAC | FโA |
| GCAG | GAG | FโA |
| $`{}_{.}{}^{.}(1,0)_{}^{6}(\frac{1}{2},\frac{1}{2})^4`$ | | |
| CCAG | CAG | FโA |
| $`{}_{.}{}^{.}(1,1)_{}^{9}(\frac{3}{2},\frac{3}{2})`$ | | |
| ACAC | AAC | AโA |
| ACAU | AAU | AโA |
| ACAG | AAG | AโA |
| ACAA | AAA | AโA |
| $`{}_{.}{}^{.}(1,0)_{}^{6}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| UCAG | UAG | AโA |
| UCAA | UAA | AโA |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,2)_{}^{3}(\frac{3}{2},\frac{3}{2})`$ | | |
| UCUC | UUC | AโF |
| UCUU | UUU | AโF |
| ACUC | AUC | AโF |
| ACUU | AUU | AโF |
| $`{}_{.}{}^{.}(0,2)_{}^{2}(\frac{1}{2},\frac{3}{2})^2`$ | | |
| CCUU | CUU | AโF |
| GCUU | GUU | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{8}(\frac{3}{2},\frac{1}{2})^1`$ | | |
| UCUG | UUG | AโF |
| UCUA | UUA | AโF |
| ACUG | AUG | AโF |
| ACUA | AUA | AโF |
| $`{}_{.}{}^{.}(0,1)_{}^{5}(\frac{1}{2},\frac{1}{2})^3`$ | | |
| CCUA | CUA | AโF |
| GCUA | GUA | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{9}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| UCAC | UAC | AโF |
| UCAU | UAU | AโF |
| $`{}_{.}{}^{.}(0,1)_{}^{6}(\frac{1}{2},\frac{1}{2})^4`$ | | |
| CCAU | CAU | AโF |
| $`{}_{.}{}^{.}(0,0)_{}^{4}(\frac{1}{2},\frac{1}{2})^4`$ | | |
| CCAA | CAA | AโA |
| $`{}_{.}{}^{.}(0,1)_{}^{6}(\frac{1}{2},\frac{3}{2})^2`$ | | |
| GCAU | GAU | AโA |
| GCAA | GAA | AโA |
So we remark:
* The deletion of C, allowed or horizontally forbidden, may happen only if it is followed by a purine. In the allowed cases, it must be followed by the nucleotide A, in agreement with the observed data.
* A nucleotide A before the deleted nucleotide C appears only in the transition $`(1,1)^9(\frac{3}{2},\frac{3}{2})`$. This feature is present in the observed data with a very low occurrence, which in our language would mean that the matrix element of $`\tau `$ between these two irreducible representations is small.
Now we consider the case of deletion of U. A detailed analysis shows that only the following deletions may happen:
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,2)_{}^{1}(\frac{1}{2},\frac{3}{2})^2`$ | | |
| CUUC | CUC | AโF |
| CUUU | CUU | AโF |
| GUUC | GUC | AโF |
| GUUU | GUU | AโF |
| $`{}_{.}{}^{.}(1,2)_{}^{2}(\frac{3}{2},\frac{3}{2})`$ | | |
| CUCC | CCC | AโF |
| GUCC | GCC | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{3}(\frac{1}{2},\frac{1}{2})^4`$ | | |
| CUAC | CAC | AโF |
| CUAU | CAU | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{6}(\frac{1}{2},\frac{1}{2})^3`$ | | |
| UUCA | UCA | AโF |
| AUCA | ACA | AโF |
| $`{}_{.}{}^{.}(2,1)_{}^{2}(\frac{3}{2},\frac{1}{2})^1`$ | | |
| UUCG | UCG | AโF |
| UUUG | UUG | AโF |
| UUUA | UUA | AโF |
| AUCG | ACG | AโF |
| AUUG | AUG | AโF |
| AUUA | AUA | AโF |
| $`{}_{.}{}^{.}(2,1)_{}^{3}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| UUGC | UGC | AโF |
| UUAC | UAC | AโF |
| UUAU | UAU | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{7}(\frac{1}{2},\frac{1}{2})^3`$ | | |
| UUGU | UGU | AโF |
| $`{}_{.}{}^{.}(0,1)_{}^{4}(\frac{1}{2},\frac{1}{2})^2`$ | | |
| CUGU | CGU | AโF |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,1)_{}^{2}(\frac{1}{2},\frac{1}{2})^3`$ | | |
| CUUG | CUG | AโF |
| GUUG | GUG | AโF |
| CUUA | CUA | AโF |
| GUUA | GUA | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{2}(\frac{3}{2},\frac{1}{2})^1`$ | | |
| CUCG | CCG | AโF |
| GUCG | GCG | AโF |
| $`{}_{.}{}^{.}(1,2)_{}^{2}(\frac{1}{2},\frac{3}{2})^1`$ | | |
| UUCU | UCU | AโF |
| AUCU | ACU | AโF |
| $`{}_{.}{}^{.}(0,2)_{}^{1}(\frac{1}{2},\frac{3}{2})^1`$ | | |
| CUCU | CCU | AโF |
| GUCU | GCU | AโF |
| $`{}_{.}{}^{.}(2,2)(\frac{3}{2},\frac{3}{2})`$ | | |
| UUCC | UCC | AโF |
| UUUC | UUC | AโF |
| UUUU | UUU | AโF |
| AUCC | ACC | AโF |
| AUUC | AUC | AโF |
| AUUU | AUU | AโF |
| $`{}_{.}{}^{.}(0,1)_{}^{3}(\frac{1}{2},\frac{1}{2})^1`$ | | |
| CUCA | CCA | AโF |
| GUCA | GCA | AโF |
| $`{}_{.}{}^{.}(1,1)_{}^{3}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| CUGC | CGC | AโF |
| $`{}_{}{}^{.}{}_{.}{}^{}`$ | | |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,1)_{}^{7}(\frac{1}{2},\frac{3}{2})^1`$ | | |
| AUGU | AGU | AโA |
| AUGA | AGA | AโA |
| $`{}_{.}{}^{.}(0,1)_{}^{4}(\frac{1}{2},\frac{3}{2})^1`$ | | |
| GUGU | GGU | AโA |
| GUGA | GGA | AโA |
| $`{}_{.}{}^{.}(1,1)_{}^{3}(\frac{1}{2},\frac{3}{2})^2`$ | | |
| GUAC | GAC | AโA |
| GUAU | GAU | AโA |
| GUAG | GAG | AโA |
| GUAA | GAA | AโA |
| $`{}_{.}{}^{.}(2,0)_{}^{2}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| UUGG | UGG | AโA |
| UUAG | UAG | AโA |
| UUAA | UAA | AโA |
| $`{}_{.}{}^{.}(1,0)_{}^{2}(\frac{3}{2},\frac{1}{2})^2`$ | | |
| CUGG | CGG | AโA |
| <sub>.</sub> Action of $`\tau _{1/2,H}^{1/2}\tau _{1/2,V}^{1/2}`$ | | |
| --- | --- | --- |
| $`{}_{.}{}^{.}(1,1)_{}^{3}(\frac{3}{2},\frac{3}{2})`$ | | |
| GUGC | GGC | AโA |
| GUGG | GGG | AโA |
| $`{}_{.}{}^{.}(1,0)_{}^{2}(\frac{1}{2},\frac{1}{2})^4`$ | | |
| CUAG | CAG | AโA |
| CUAA | CAA | AโA |
| $`{}_{.}{}^{.}(2,1)_{}^{3}(\frac{3}{2},\frac{3}{2})`$ | | |
| AUGC | AGC | AโA |
| AUAC | AAC | AโA |
| AUAU | AAU | AโA |
| AUGG | AGG | AโA |
| AUAG | AAG | AโA |
| AUAA | AAA | AโA |
| $`{}_{.}{}^{.}(0,0)_{}^{2}(\frac{1}{2},\frac{1}{2})^2`$ | | |
| CUGA | CGA | AโA |
| $`{}_{}{}^{.}{}_{.}{}^{}`$ | | |
So we remark:
* The deletion of U may happen only if it is followed by A or by G. In the observed data only A is considered; however in the reported deletion of U are about 1/4 with respect to the reported deletion of C. So our modelisation just foresees a different environment for the deletion of U and C.
* The last nucleotide in the four-nucleotide sequence in which the deletion occurs may be any nucleotide, but the case in which it is a purine seems more frequent than the case in which it is a pyrimidine.
* There are no transition which are only horizontally forbidden.
In conclusion, both from considering the transitions on the $`K`$-chains ($`K=2,3`$) to the $`(K1)`$-chains or the transition from the four-nucleotide states to the three-nucleotide states under the action of the crystal tensor operators, we deduce that the deletion of a pyrimidine may happen if it is followed by a purine. In particular, for the deletion of C the preferred purine is the adenine A, whilst for the deletion of U also the guanine G may appear. This makes a difference between the two cases and it would be extremely interesting to see if more accurate data may confirm this asymmetry. Moreover the next following nucleotide may be of any type but there is indication that a purine is preferred. So our mathematical scheme explains the main features of the observed data . A more quantitative analysis should require higher statistics in the experimental data.
## 7 Recent theoretical approaches: a comparison
The use of continuous symmetries in the genetic code has been considered by different teams these recent years<sup>2</sup><sup>2</sup>2See section โSymmetry techniques in Biological Systemsโ in Proc. XXII Int. Coll. on Group Theoretical Methods in Physics, pp. 142-165.. It appears of some importance to summarize each of these approaches, and to make clear how the model we propose differ from them.
In 1993, an underlying symmetry based on a continuous group has been proposed . More precisely, considering the eukaryotic code, the authors tried to answer the following question: is it possible to determine a Lie algebra $`๐ข`$ carrying a 64-dimensional irreducible representation $`R`$ and admitting a subalgebra $``$ such that the decomposition of $`R`$ into irreducible multiplets under $``$ gives exactly the 21 different multiplets, the different codons in each of the first 20 multiplets being associated to the same amino-acid, the last multiplet containing the stop codons ? They proposed as starting symmetry the symplectic algebra $`sp(6)`$, which indeed admits an irreducible representation of dimension 64, equal to the number of different codons, with the successive breakings:
$$sp(6)sp(4)su(2)su(2)su(2)su(2)su(2)U(1)su(2)su(2)U(1)U(1)$$
(67)
Such a chain of symmetry breaking could be considered as reflecting the evolution of the genetic code, the six amino-acids relative to the codons in the irreducible representations obtained after the first breaking (in which 64 = 16 + 4 + 20 + 10 + 12 + 2) appearing as primordial amino-acids in their approach. However, the authors were obliged, in order to reproduce the actual multiplet pattern, to assume in the final breaking, a partial breaking or a โfreezingโ in the sense that the breaking of the last $`su(2)`$ into $`U(1)`$ does not occur for all the multiplets. As an example, such a freezing has to be imposed to the sextets corresponding to Leu and Ser, which otherwise would decompose into three doublets. In the same way, freezing will forbid the doublets related to Lys and Cys to split into singlets.
In a second further paper, dated 1997 , a refinement of this approach has been considered, with the use of Lie groups instead of Lie algebras: then, global properties, for example non connexity of $`O(2)=U(1)\times _2`$, can be exploited. In this context, the authors proposed another chain of breaking starting with the exceptional group $`G_2`$, which also allows a 64 dimensional irreducible representation. But here again, the freezing pathology cannot be avoided.
One can also mention the work of where the unifying algebra before breaking is $`so(14)`$.
Meantime (1997), interpreting the double origin of the nucleotides, each arising either from purine or from pyrimidine, as a $`_2`$-grading a supersymmetric model was proposed , involving superalgebras for such a program. The $`_2`$-grading specific of a simple superalgebra is there used to separate purine and pyrimidine: indeed, by putting the four nucleotids in the the 4 dimensional representation of $`su(2/1)`$ one can confer to the A and G purines (R) an even grading, and to the C and U pyrimidines (Y) an odd grading; note that the R states are then in the $`su(2)`$ doublet and the Y ones $`su(2)`$ singlets. The notion of polarity spin is also introduced, allowing to distinguish the C and G nucleotides with two locally polarized sites, from the A and U ones with three polarized sites: the C and G (resp. A and U) will be assigned in a doublet (resp. in singlets) of another $`su(2)`$. Then the authors consider the sum of algebras: $`su(2)su(2)su(2|1)`$ with the first (second) $`su(2)`$ acting as polarity spin on the first (second) nucleotid of a codon, and the $`su(2|1)`$ acting on the third nucleotid only. Moreover the two $`su(2)`$ would act in an alternating way on the first and second position, that is as $`1/2`$, $`1/2`$ and $`1/2`$, $`1/2`$. This sum of algebras can be embedded in the superalgebra $`su(6|1)`$, which admits a 64 dimensional irreducible representation, and could be also used for a superalgebraic approach to the genetic code evolution, with the chain of symmetry breaking:
$`su(6|1)su(2)su(3|1)su(2)su(2)su(2|1)U(1)U(1)su(2|1)`$ (68)
$`U(1)U(1)gl(1|1)`$
Again the problem of freezing, that is the last breaking applies to some but not all the multiplets, is present with this choice of (super)algebras.
It seems necessary to remark that in this proposal which implies (super)algebras acting in the same time on nucleotides and on codons โ one must say in a rather complicated way โ the nucleotides cannot appear as building blocks from which one algebraically constructs the codons, by performing tensorial products of representations, as is the case of our model. In fact, the problem of ordering the nucleotides inside a codon forbids this natural way of proceeding as long as only usual (super)algebras are involved. Note that it is the limit of quantum algebras that we use in our approach: then, we have at hand the so-called crystal bases, which exactly solve the ordering problem.
In a last month preprint, two authors of the same team proposed to fit biophysical properties of nucleic acids by constructing polynomials in 6 coordinates in the 64 dimensional codon space. As already mentioned in sect. 4, the two coordinates they associate to each nucleotide is direcetly related to the nucleotide eigenvalues of our model. The authors present their computations as independent of a particular choice of algebra or superalgebra as long as the underlying algebra is of rank 6 โ which is in particular the dimension of the Cartan subalgebra of $`su(6|1)`$ โ and admits a 64 dimensional irreducible representation. We note that our model does allow to calculate the biophysical quantities considered in ref. without the constraint on representations, but more importantly, with only a two rank algebra.
A detailed and systematic study of superalgebras and superalgebra breaking chains has been performed by the authors of : it is the orthosymplectic $`osp(5|2)`$ superalgebra which emerges from their algebraic analysis.
Finally, it is amazing to remark that, just a few years after the the concept of genetic code was formulated, an attempt to give a mathematical description of its properties was started by the russian physicist Yu. B. Rumer . Indeed he remarked that the 16 *roots*, i.e. the combinations of the first two codons, divide in a *strong octet* which form quartets ou sub-part of sextets and a *weak octet* which form doublets, triplets and singlets, attempting to give a systematic description of the genetic code. A few years after, with B.G. Konopelโchenko they formulated the strong assumption that with respect to any property of the codons the 16 roots can be gathered into two octets with opposite โchargeโ, whose positive (negative) value respectively characterizes the strong and weak roots. This description comes out naturally in our model, such a charge $`Q`$ being defined in Eq. (5) of sect. 2.
## 8 Conclusion
Our model is based on the algebra $`๐ฐ_{q0}(sl(2)sl(2))`$ that we have chosen for two main characteristics. First it encodes the stereochemical property of a base, and also reflects the complementarity rule, by conferring quantum numbers to each nucleotide. Secondly, it admits representation spaces or crystal bases in which an ordered sequence of nucleotides or codon can be suitably characterized. Let us emphasize that $`๐ฐ_{q0}(sl(2)sl(2))`$ is really neither a Lie algebra nor an enveloping deformed algebra. We still use in a loose sense the word algebra, just to emphasize the fact that we use largely the mathematical tools of representation space, tensor operators etc. which are typical of the algebraic structures. Let us add that it is a remarkable property of a quantum algebra in the limit $`q0`$ to admit representations, obtained from the tensorial product of basic ones, in which each state appears as a unique sequence of ordered basic elements.
In this framework, the correspondence codon/amino-acid is realized by the operator $`_c`$, constructed out of the symmetry algebra, and acting on codons: the eigenvalues provided by $`_c`$ on two codons will be equal or different depending on whether the two codons are associated to the same or to two different amino-acids. It is remarkable that this correspondence can be obtained for all the genetic codes and that the reading operators have a bulk common to the various genetic codes (the prototype reading operator) and differ only for a few additive terms, analogous to perturbative terms present in most Hamiltonians describing complex physical systems. Moreover they depend on parameters, presently assumed as constants, which in principle can be considered as functions of suitable variables. These feature may be of some interest in the study of the evolution of the genetic code, problem which has not yet been tackled in our model.
Then, restricting to the case of states made of two nucleotides, the experimental values of the free energy, released by base pairing in the formation of double stranded nucleic acids, of the hydrophibicity and of the hydrophilicity have been fitted with expressions depending respectively on 2, 4 and 4 parameters and constructed out of the generators of $`๐ฐ_{q0}(sl(2)sl(2))`$.
The model does not necessarily assign the codons in a multiplet (in particular the quartets, sextets and triplet) to the same irreducible representation. Let us remark that the assignments of the codons to the different irreducible representations is a straightforward consequence of the tensor product, once assigned the nucleotides to the fundamental irreducible representation. This feature is relevant, since it can explain the correlation between the branching ratios of the codon usage of different codons coding the same amino-acid as discussed in and . Here we have shown that the universal pattern (inside the class of vertebrates) of $`B_{AG}/B_{UC}`$ can simply be reproduced in our model.
Moreover our mathematical description of the genetic code allows a modelisation of some biological process. A first step in this direction has been presented in sect. 6, where we have shown that the observed data related to the a pyrimidine deletion can be simulated by introducing the concept of $`q0`$ โ or crystal โ tensor operator. Finally let us mention some directions for future development of our model. Going further in the analysis of the branching ratios, we want to refine our analysis and make a more detailed study taking into account the dependence on the family of biological species. Indeed preliminary analysis on plants, invertebrates and bacteriae shows that, even if the pattern of the correlation is still approximatively present, large deviations appear which presumably exhibit evidence that the dependence on subclass or family of biological species cannot any more be neglected, differently to the case of vertebrates. A further investigation of the possibility of mathematically modelising or simulating biological processes, in particular mutations, by crystal tensor operators, is in progress. Other questions are still to be investigated: in particular how could the genetic code evolution be reproduced in our model ?
Figure 1 (continued)
$`(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})`$ $`=`$ $`(\frac{1}{2},\frac{1}{2})\left[(\frac{3}{2},\frac{3}{2})2(\frac{3}{2},\frac{1}{2})2(\frac{1}{2},\frac{3}{2})4(\frac{1}{2},\frac{1}{2})\right]`$
$`=`$ $`(2,2)\mathrm{\hspace{0.17em}3}(2,1)\mathrm{\hspace{0.17em}3}(1,2)\mathrm{\hspace{0.17em}9}(1,1)\mathrm{\hspace{0.17em}2}(2,0)`$
$`\mathrm{\hspace{0.17em}2}(0,2)\mathrm{\hspace{0.17em}6}(1,0)\mathrm{\hspace{0.17em}6}(0,1)\mathrm{\hspace{0.17em}4}(0,0)`$
One has (The upper label denotes different irreducible representations):
$$(\frac{1}{2},\frac{1}{2})(\frac{3}{2},\frac{3}{2})=(2,2)(2,1)^1(1,2)^1(1,1)^1$$
where
$$\begin{array}{ccc}(2,2)=& & (1,2)^1=\\ \left(\begin{array}{ccccc}\text{CCCC}& \text{UCCC}& \text{UUCC}& \text{UUUC}& \text{UUUU}\\ \text{GCCC}& \text{ACCC}& \text{AUCC}& \text{AUUC}& \text{AUUU}\\ \text{GGCC}& \text{AGCC}& \text{AACC}& \text{AAUC}& \text{AAUU}\\ \text{GGGC}& \text{AGGC}& \text{AAGC}& \text{AAAC}& \text{AAAU}\\ \text{GGGG}& \text{AGGG}& \text{AAGG}& \text{AAAG}& \text{AAAA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CUCC}& \text{CUUC}& \text{CUUU}& & \\ \text{GUCC}& \text{GUUC}& \text{GUUU}& & \\ \text{GACC}& \text{GAUC}& \text{GAUU}& & \\ \text{GAGC}& \text{GAAC}& \text{GAAU}& & \\ \text{GAGG}& \text{GAAG}& \text{GAAA}& & \end{array}\right)\\ & & \\ (2,1)^1=& & (1,1)^1=\\ \left(\begin{array}{ccccc}\text{CGCC}& \text{UGCC}& \text{UACC}& \text{UAUC}& \text{UAUU}\\ \text{CGGC}& \text{UGGC}& \text{UAGC}& \text{UAAC}& \text{UAAU}\\ \text{CGGG}& \text{UGGG}& \text{UAGG}& \text{UAAG}& \text{UAAA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CACC}& \text{CAUC}& \text{CAUU}& & \\ \text{CAGC}& \text{CAAC}& \text{CAAU}& & \\ \text{CAGG}& \text{CAAG}& \text{CAAA}& & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{3}{2},\frac{1}{2})^1=(2,1)^2(2,0)^1(1,1)^2(1,0)^1$$
where
$$\begin{array}{ccc}(2,1)^2=& & (1,1)^2=\\ \left(\begin{array}{ccccc}\text{CCCG}& \text{UCCG}& \text{UUCG}& \text{UUUG}& \text{UUUA}\\ \text{GCCG}& \text{ACCG}& \text{AUCG}& \text{AUUG}& \text{AUUA}\\ \text{GGCG}& \text{AGCG}& \text{AACG}& \text{AAUG}& \text{AAUA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CUCG}& \text{CUUG}& \text{CUUA}& & \\ \text{GUCG}& \text{GUUG}& \text{GUUA}& & \\ \text{GACG}& \text{GAUG}& \text{GAUA}& & \end{array}\right)\\ & & \\ (2,0)^1=& & (1,0)^1=\\ \left(\begin{array}{ccccc}\text{CGCG}& \text{UGCG}& \text{UACG}& \text{UAUG}& \text{UAUA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CACG}& \text{CAUG}& \text{CAUA}& & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{3}{2},\frac{1}{2})^2=(2,1)^3(2,0)^2(1,1)^3(1,0)^2$$
where
$$\begin{array}{ccc}(2,1)^3=& & (1,1)^3=\\ \left(\begin{array}{ccccc}\text{CCGC}& \text{UCGC}& \text{UUGC}& \text{UUAC}& \text{UUAU}\\ \text{GCGC}& \text{ACGC}& \text{AUGC}& \text{AUAC}& \text{AUAU}\\ \text{GCGG}& \text{ACGG}& \text{AUGG}& \text{AUAG}& \text{AUAA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CUGC}& \text{CUAC}& \text{CUAU}& & \\ \text{GUGC}& \text{GUAC}& \text{GUAU}& & \\ \text{GUGG}& \text{GUAG}& \text{GUAA}& & \end{array}\right)\\ & & \\ (2,0)^2=& & (1,0)^2=\\ \left(\begin{array}{ccccc}\text{CCGG}& \text{UCGG}& \text{UUGG}& \text{UUAG}& \text{UUAA}\end{array}\right)& & \left(\begin{array}{ccccc}\text{CUGG}& \text{CUAG}& \text{CUAA}& & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{3}{2})^1=(1,2)^2(0,2)^1(1,1)^4(0,1)^1$$
where
$$\begin{array}{ccc}(1,2)^2=\left(\begin{array}{ccccc}\text{CCCU}& \text{UCCU}& \text{UUCU}& & \\ \text{GCCU}& \text{ACCU}& \text{AUCU}& & \\ \text{GGCU}& \text{AGCU}& \text{AACU}& & \\ \text{GGGU}& \text{AGGU}& \text{AAGU}& & \\ \text{GGGA}& \text{AGGA}& \text{AAGA}& & \end{array}\right)& & (0,2)^1=\left(\begin{array}{ccccc}\text{CUCU}& & & & \\ \text{GUCU}& & & & \\ \text{GACU}& & & & \\ \text{GAGU}& & & & \\ \text{GAGA}& & & & \end{array}\right)\\ & & \\ (1,1)^4=\left(\begin{array}{ccccc}\text{CGCU}& \text{UGCU}& \text{UACU}& & \\ \text{CGGU}& \text{UGGU}& \text{UAGU}& & \\ \text{CGGA}& \text{UGGA}& \text{UAGA}& & \end{array}\right)& & (0,1)^1=\left(\begin{array}{ccccc}\text{CACU}& & & & \\ \text{CAGU}& & & & \\ \text{CAGA}& & & & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{3}{2})^2=(1,2)^3(0,2)^2(1,1)^5(0,1)^2$$
where
$$\begin{array}{ccc}(1,2)^3=\left(\begin{array}{ccccc}\text{CCUC}& \text{UCUC}& \text{UCUU}& & \\ \text{GCUC}& \text{ACUC}& \text{ACUU}& & \\ \text{GGUC}& \text{AGUC}& \text{AGUU}& & \\ \text{GGAC}& \text{AGAC}& \text{AGAU}& & \\ \text{GGAG}& \text{AGAG}& \text{AGAA}& & \end{array}\right)& & (0,2)^2=\left(\begin{array}{ccccc}\text{CCUU}& & & & \\ \text{GCUU}& & & & \\ \text{GGUU}& & & & \\ \text{GGAU}& & & & \\ \text{GGAA}& & & & \end{array}\right)\\ & & \\ (1,1)^5=\left(\begin{array}{ccccc}\text{CGUC}& \text{UGUC}& \text{UGUU}& & \\ \text{CGAC}& \text{UGAC}& \text{UGAU}& & \\ \text{CGAG}& \text{UGAG}& \text{UGAA}& & \end{array}\right)& & (0,1)^2=\left(\begin{array}{ccccc}\text{CGUU}& & & & \\ \text{CGAU}& & & & \\ \text{CGAA}& & & & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})^1=(1,1)^6(1,0)^3(0,1)^3(0,0)^1$$
where
$$\begin{array}{ccc}(1,1)^6=\left(\begin{array}{ccccc}\text{CCCA}& \text{UCCA}& \text{UUCA}& & \\ \text{GCCA}& \text{ACCA}& \text{AUCA}& & \\ \text{GGCA}& \text{AGCA}& \text{AACA}& & \end{array}\right)& & (0,1)^3=\left(\begin{array}{ccccc}\text{CUCA}& & & & \\ \text{GUCA}& & & & \\ \text{GACA}& & & & \end{array}\right)\\ & & \\ (1,0)^3=\left(\begin{array}{ccccc}\text{CGCA}& \text{UGCA}& \text{UACA}& & \end{array}\right)& & (0,0)^1=\left(\begin{array}{ccccc}\text{CACA}& & & & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})^2=(1,1)^7(1,0)^4(0,1)^4(0,0)^2$$
where
$$\begin{array}{ccc}(1,1)^7=\left(\begin{array}{ccccc}\text{CCGU}& \text{UCGU}& \text{UUGU}& & \\ \text{GCGU}& \text{ACGU}& \text{AUGU}& & \\ \text{GCGA}& \text{ACGA}& \text{AUGA}& & \end{array}\right)& & (0,1)^4=\left(\begin{array}{ccccc}\text{CUGU}& & & & \\ \text{GUGU}& & & & \\ \text{GUGA}& & & & \end{array}\right)\\ & & \\ (1,0)^4=\left(\begin{array}{ccccc}\text{CCGA}& \text{UCGA}& \text{UUGA}& & \end{array}\right)& & (0,0)^2=\left(\begin{array}{ccccc}\text{CUGA}& & & & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})^3=(1,1)^8(1,0)^5(0,1)^5(0,0)^3$$
where
$$\begin{array}{ccc}(1,1)^8=\left(\begin{array}{ccccc}\text{CCUG}& \text{UCUG}& \text{UCUA}& & \\ \text{GCUG}& \text{ACUG}& \text{ACUA}& & \\ \text{GGUG}& \text{AGUG}& \text{AGUA}& & \end{array}\right)& & (0,1)^5=\left(\begin{array}{ccccc}\text{CCUA}& & & & \\ \text{GCUA}& & & & \\ \text{GGUA}& & & & \end{array}\right)\\ & & \\ (1,0)^5=\left(\begin{array}{ccccc}\text{CGUG}& \text{UGUG}& \text{UGUA}& & \end{array}\right)& & (0,0)^3=\left(\begin{array}{ccccc}\text{CGUA}& & & & \end{array}\right)\end{array}$$
$$(\frac{1}{2},\frac{1}{2})(\frac{1}{2},\frac{1}{2})^4=(1,1)^9(1,0)^6(0,1)^6(0,0)^4$$
where
$$\begin{array}{ccc}(1,1)^9=\left(\begin{array}{ccccc}\text{CCAC}& \text{UCAC}& \text{UCAU}& & \\ \text{GCAC}& \text{ACAC}& \text{ACAU}& & \\ \text{GCAG}& \text{ACAG}& \text{ACAA}& & \end{array}\right)& & (0,1)^6=\left(\begin{array}{ccccc}\text{CCAU}& & & & \\ \text{GCAU}& & & & \\ \text{GCAA}& & & & \end{array}\right)\\ & & \\ (1,0)^6=\left(\begin{array}{ccccc}\text{CCAG}& \text{UCAG}& \text{UCAA}& & \end{array}\right)& & (0,0)^4=\left(\begin{array}{ccccc}\text{CCAA}& & & & \end{array}\right)\end{array}$$ |
warning/0003/astro-ph0003234.html | ar5iv | text | # Blue compact galaxies and the primordial 4Helium abundance
## 1. Introduction
Blue compact galaxies (BCG) are low-luminosity (M<sub>B</sub> $`>`$ -18) systems which are undergoing an intense burst of star formation in a very compact region (less than 1 kpc) which dominates the light of the galaxy (Figure 1) and which shows blue colors and a HII region-like emission-line optical spectrum (Figure 2). BCGs are ideal laboratories in which to measure the primordial <sup>4</sup>Helium abundance because of several reasons:
1) With an oxygen abundance O/H ranging between 1/50 and 1/3 that of the Sun, BCGs are among the most metal-deficient gas-rich galaxies known. Their gas has not been processed through many generations of stars, and thus best approximates the pristine primordial gas. Izotov & Thuan (1999) have argued that BCGs with O/H less than $``$ 1/20 that of the Sun may be genuine young galaxies. Their argument is based on the observed constancy and very small scatter of the C/O and N/O ratios in extremely metal-deficient BCGs with 12 + log O/H $`<`$ 7.6, which they interpret as implying that the C and N in these galaxies have been made in the same massive stars (M $`>`$ 9 M) which manufactured O, and that intermediate-mass stars (3 M $`<`$ M $`<`$ 9 M) have not had time to release their nucleosynthetic products. Since the main-sequence lifetime of a 9 M star is $``$ 40 Myr, Izotov & Thuan (1999) suggest that very metal-deficient BCGs are younger than $``$ 100 Myr. Thus the primordial Helium mass fraction Y<sub>p</sub> can be derived accurately in very metal-deficient BCGs with only a small correction for Helium made in stars.
2) Because of the relative insensitivity of <sup>4</sup>He production to the baryonic density of matter, Y<sub>p</sub> needs to be determined to a precision better than 5% to provide useful cosmological constraints. This precision can in principle be achieved by using BCGs because their optical spectra show several He I recombination emission lines and very high signal-to-noise ratio emission-line spectra with moderate spectral resolution of BCGs can be obtained at large telescospes (4 m class or larger) coupled with efficient and linear CCD detectors with a relatively modest investment of telescope time. The theory of nebular emission is well understood and the theoretical He I recombination coefficients calculated by Brocklehurst (1972) and Smits (1996) are well known enough to allow to convert He emission-line strengths into abundances with the desired accuracy.
## 2. The primordial He abundance from extrapolation of the Y-O/H and Y-N/H linear regressions
### 2.1. A new large sample of Blue Compact Galaxies
Peimbert & Torres-Peimbert (1974, 1976) first noted the correlation between He and O abundances in a small sample of dwarf magellanic irregulars and BCGs, and they proposed to determine Y<sub>p</sub> by linear extrapolation of the correlation to O/H = 0. Later, Pagel, Terlevich & Melnick (1986) proposed to use also the Y-N/H correlation for the determination of Y<sub>p</sub>, to take into account the temporary local excess of helium and nitrogen due to pollution by winds from massive stars. Many attempts at determining Y<sub>p</sub> have been made, using the Y versus O/H and Y versus N/H correlations on various samples of dwarf irregulars and BCGs (e.g. Pagel et al. 1992, Izotov, Thuan & Lipovetsky (1994, 1997, hereafter ITL94 and ITL97; Olive, Steigman & SKillman 1997, hereafter OSS97; Izotov & Thuan 1998ab, hereafter IT98ab).
Before our work, the largest, most accurate and consistent published data set was by Pagel et al. (1992). Their observations were reduced in a uniform manner and they paid careful attention to such points as the correction for the unseen neutral helium and electron collisional effects which may make some He I lines deviate from their recombination values. Pagel et al. (1992) obtained Y<sub>p</sub> = 0.228$`\pm `$0.005, below the limit set by the standard hot big bang model of nucleosynthesis (SBBN) and consistent with it only at the 2$`\sigma `$ level. This prompted us to consider obtaining another measurement of Y<sub>p</sub> from an independent data set with as high or better precision to test SBBN.
Starting in 1993, we embarked on a large-scale program to obtain high signal-to-noise ratio spectra for a relatively large sample of BCGs assembled from several objective-prism surveys: the First Byurakan or Markarian survey (Markarian et al. 1989), the Second Byurakan Survey (SBS, Izotov et al. 1993) and the University of Michigan survey (Salzer, MacAlpine & Boroson 1989). The SBS sample was particularly interesting because it contained about a dozen BCGs with O/H less than 1/15 of (O/H), more than doubling the number of such known low-metallicity BCGs. The total sample consists of 45 H II regions in 42 BCGs. The data have been published in a series of papers in the Astrophysical Journal: ITL94, ITL97 and IT98ab.
### 2.2. Methodology
There are a number of features which distinguish our work from previous efforts in determining the primordial He abundance. Our methodology is described in detail in ITL94, ITL97, and IT98ab.
1) We have observed all the galaxies in our sample with the same telescopes (the Kitt Peak 4 m and 2.1 m telescopes) and instrumental set up, and the data were all reduced in a homogeneous way. This differs from OSS97, for example, which used a more heterogeneous sample of BCGs observed by different observers on different telescopes, with some of the data obtained many years ago with nonlinear detectors. A uniform sample is essential to minimize as much as possible the artificial scatter introduced by assembling different data sets reduced in different ways.
2) To derive the He mass fraction, previous authors use mainly one He emission line, He I 6678. Correction to this lineโs intensity is usually made only for one effect, electron collisional enhancement. This correction is usually carried out adopting the electron density derived from the \[S II\] 6717/6731 emission-line ratio. The approach just described has several shortcomings. The metastability of the 2<sup>3</sup>S state of He I can also lead to possible radiative transfer effects (also called fluorescence effects) in the triplet lines which may be enhanced at the expense of the He I 3889 line (Robbins 1968). When a single He I emission line is used, one cannot distinguish between between electron collisional and radiative transfer effects. Thus, fluorescent enhancement is neglected, while it may be important. Furthermore, at the low electron number densities N<sub>e</sub> which often characterize the HII regions in BCGs, the determination of N<sub>e</sub> from \[S II\] emission lines is very uncertain. In the majority of cases, N<sub>e</sub> is arbitrarily set to 100 cm<sup>-3</sup>. This assumption can lead to artificially low He abundance, as in the case of the southeast component of I Zw 18 where the true N<sub>e</sub> is only $``$ 10 cm<sup>-3</sup> (Izotov et al. 1999). More importantly, setting N<sub>e</sub>(He II) equal to N<sub>e</sub>(S II) is not physically reasonable as the S<sup>+</sup> and He<sup>+</sup> regions are not expected to coincide, given the large difference in the S I and He I ionization potentials.
To remedy these problems, we have proposed a self-consistent method in which we use all five brightest He I emission-lines in the optical range (the He I 3889, 4471, 5876, 6678 and 7065 lines) and solve simultaneously for N<sub>e</sub>(He II) and the optical depth in the He I 3889 line so that the He I 3889/4471, 5876/4471, 6678/4471 and 7065/4471 line ratios have their recombination values, after correction for both collisional (Kingdon & Ferland 1995) and fluorescent (Robbins 1968) enhancements. The He I 3889 and 7065 lines play an important role because they are particularly sensitive to both optical depth and electron number density.
### 2.3. Underlying stellar absorption
Effects other than collisional and fluorescent enhancements can also change He I line intensities. An important effect is the underlying stellar absorption in He I lines caused by hot OB stars which decreases the intensities of nebular He I lines. This effect is most important for the emission lines with the smallest equivalent widths.
The neglect of He I underlying stellar absorption can lead to a severe underestimate of the He mass fraction. One of the most spectacular examples is that of the BCG I Zw18. This object plays a key role in the determination of the primordial He abundance because, with an O/H only 1/50 that of the Sun, it is the most metal-deficient BCG known and has great influence on the derived slopes and intercepts of the Y-O/H and Y-N/H linear regression lines. Figure 2 shows very high signal-to-noise ratio spectrophotometric observations of I Zw 18 obtained with the Multiple Mirror Telescope (MMT), with the slit oriented so as to go through the two main centers of star formation in the BCG, the so-called NW and SE components (Figure 1, left). Comparison of the spectrum of the NW component (Figure 2a) with that of the SE component (Figure 2b) shows clearly that underlying stellar absorption is much more important in the NW than in the SE component: all marked He I lines in the spectrum of the SE component are in emission while the two He I 4026 and 4921 lines are in absorption and the He I 4471 emission line is barely visible in the spectrum of the NW component. Prior to our work, most of the He work relied on measurements of Y in the NW component because of its larger brightness as compared to the SE component (Figure 1, left). This led to a systematic underestimate of Y in I Zw 18 . Izotov et al. (1999) derive the implausibly small Y(4471)= 0.169$`\pm `$0.023 and Y(5876)= 0.192$`\pm `$0.007 from the He I 4471 and 5876 lines respectively. In addition to underlying stellar absorption, the 5876 line intensity in I Zw 18 is reduced further by absorption from the Galactic interstellar 5890 and 5896 Na I lines.
While the NW component cannot be used for He determination, the SE component is better suited as the influence of underlying stellar absorption on the He I emission line intensities is significantly smaller in this component. This can be seen in the left panel of Figure 3 which shows the spatial distributions of the He I nebular emission-line equivalent widths in I Zw 18: they are systematically larger in the SE component than in the NW one, implying less absorption.
### 2.4. Results
Figure 4 shows the Y-O/H and Y-N/H linear regressions for the whole sample of 45 H II regions in BCGs. The sample includes most of the very metal-deficient BCGs known, including the two most extreme ones, the SE component of I Zw 18 and SBS 0335โ052 with O/H about 1/43 that of the Sun. We obtain Y<sub>p</sub>= 0.2443$`\pm `$0.0015 with dY/dZ = 2.4$`\pm `$1.0 (IT98b). Our Y<sub>p</sub> is considerably higher than those derived by other groups which range from 0.228$`\pm `$0.005 (Pagel et al. 1992) to 0.234$`\pm `$0.002 (OSS97). At the same time, our derived slope is significantly smaller than those of other authors, dY/dZ = 6.7$`\pm `$2.3 for Pagel et al. (1992) and dY/dZ = 6.9$`\pm `$1.5 for OSS97. This shallower slope is in good agreement with the value derived from stellar data for the Milky Wayโs disk and with simple models of galactic evolution of BCGs with well-mixed homogeneous outflows.
## 3. He abundance in the two most metal-deficient blue compact galaxies known
### 3.1. I Zw 18 and SBS 0335โ052
Instead of the statistical approach described above, we can also derive the primordial He abundance from accurate measurements of the He abundance in a few objects selected to have very low O/H to minimize the amount of He manufactured in stars.
Izotov et al. (1999) have carried out such a study for the two most metal-deficient BCGs known. I Zw 18 and SBS 0335โ052 provide a study in contrast concerning the different physical mechanisms which may modify the He I emission-line intensities. While in I Zw 18, the electron number density is small (N<sub>e</sub> $`<`$ 100 cm<sup>-3</sup>) and collisional enhancement has a minor effect on the derived helium abundance, N<sub>e</sub> is much higher in SBS 0335โ052 (N<sub>e</sub> $``$ 500 cm<sup>-3</sup> in the central part of the H II region). Additionally, the linear size of the H II region in SBS 0335โ052 is $``$ 5 times larger than in I Zw 18, suggesting that it may be optically thick for some He I transitions. In fact, both collisional and fluorescent enhancements of He I emission lines play an important role in this galaxy. By contrast, underlying stellar He I absorption is much less important in SBS 0335โ052 than in I Zw 18. Since the equivalent widths (EW) of the He I emission lines scale roughly as the H$`\beta `$ EWs, it is evident that this is the case from Figures 5c and 5f which show the spatial distribution of EW(H$`\beta `$) in both BCGs. In SBS 0335โ052, EW(H$`\beta `$) has a lowest value of 160 ร
and increases to $``$ 300 ร
in the outer parts, while in I Zw 18 , EW(H$`\beta `$) is only 34 ร
in the center of the NW component. Given equal EWs for He I absorption lines in both BCGs, we may expect the effect of underlying stellar absorption to be $``$ 5 times smaller in SBS 0335โ052 than in I Zw 18.
To disentangle the various effects which may make the He I emission-line intensities deviate from their recombination values, it is thus essential to use as many He I lines as possible in a self-consistent method as decribed above. An important and essential check that all corrections have been properly applied is the agreement between the He mass fraction Y derived independently from each He line. Figure 6 shows the Ys derived from the 4471, 5876 and 6678 He I emission lines in SBS 0335โ052 at different spatial locations. It is clear that the self-consistent method (Figure 6a) gives much better agreement between the different lines. The Ys derived from the 4471 line are systematically below because only collisional and fluorescence effects have been taken into account and not underlying stellar absorption, and because the 4471 line is more subject to the latter effect. By contrast, there is not very good agreement between the Ys from different lines when N<sub>e</sub>(He II) is set equal to N<sub>e</sub>(S II) and only collisional enhancement is taken into account (Figure 6b).
### 3.2. Results
Izotov et al. (1999) derive Y = 0.243$`\pm `$0.007 for the SE component of I Zw 18 in very good agreement wth the value found by IT98a and Y = 0.2463$`\pm `$0.0015 for SBS 0335โ052, excluding the He I 4471 line. The weighted mean is then Y = 0.2462$`\pm `$0.0015. Using dY/dZ = 2.4 (IT98b), the stellar He contribution is 0.0010, giving a primordial value Y<sub>p</sub> = 0.2452$`\pm `$0.0015, in excellent agreement with the value 0.2443$`\pm `$0.0015 derived from extrapolation of the Y-O/H and Y-N/H regression lines for our large BCG sample. It is, however, higher than the value Y$`p`$ = 0.2345$`\pm `$0.0030 derived by Peimbert & Peimbert (2000) from Magellanic Clouds H II regions. These authors suggest that two systematic effects may cause the disagreement: the presence of neutral hydrogen inside the helium Stromgren sphere and temperature fluctuations in our BCGs.
There is no evidence that the first effect is important in our objects. In our work, we have used the โradiation softness parameterโ of Vilchez & Pagel (1988) to estimate the correction factor for neutral helium and found the fraction of neutral helium to be insignificant ($`<`$ 2%) in all our objects, i.e their HII and He II Stromgren spheres are coincident to a very good approximation. This conclusion is corroborated by the detailed modeling of I Zw 18 by Stasinska & Schaerer (1999) who found the amount of neutral Helium to be negligible. On the other hand, they did find the observed T<sub>e</sub>(OIII) to be $``$ 15% higher than predicted by the photoionization model. Future progress in the determination of the primordial <sup>4</sup>He abundance using BCGs will rely on the discovery of more I Zw 18-like objects and on detailed modeling of very high signal-to-noise ratio and high-spectral resolution spectra of a few very metal-deficient BCGs (those with O/H less than 1/20 of solar) to look into such systematic effects as those mentioned above.
## 4. Cosmological implications
Figure 7 shows the primordial abundances of <sup>4</sup>He, D, <sup>3</sup>He, and <sup>7</sup>Li predicted by standard big bang nucleosynthesis theory as a function of the baryon-to photon number ratio $`\eta `$. The dashed lines are 1 $`\sigma `$ uncertainties in model calculations. The solid boxes show the 1 $`\sigma `$ predictions of $`\eta `$ as inferred from our derived primordial abundance of <sup>4</sup>He, and the primordial abundances of D (Levshakov, Kegel & Takahara 1999; Burles & Tytler 1998), <sup>3</sup>He (Rood et al. 1998) and <sup>7</sup>Li (Bonifacio & Molaro 1997; Vauclair & Charbonnel 1998; Pinsonneault et al. 1999). All these determinations are consistent to within 1 $`\sigma `$, although the most stringent constraint is provided by D. For comparison, we have also plotted the Y<sub>p</sub> derived by OSS97 which is low, partly because underlying stellar absorption was not taken into account in I Zw 18. Their Y<sub>p</sub> = 0.234 happens also to be the value obtained by Peimbert & Peimbert (2000) in the Magellanic Clouds. This low value would have been consistent with the primordial D abundance obtained by Songaila et al (1997) which is one order of magnitude higher than the value obtained by Burles & Tytler (1998), except that this high value is now believed to be erroneous.
Our Y<sub>p</sub> = 0.2452$`\pm `$0.0015 value implies a baryon-to-photon number ratio $`\eta `$ = 4.7$`{}_{0.8}{}^{}{}_{}{}^{+1.0}`$ $`\times `$10<sup>-10</sup>. This translates to a baryon mass fraction $`\mathrm{\Omega }_b`$$`h_{50}^2`$ = 0.068$`{}_{0.012}{}^{}{}_{}{}^{+0.015}`$ where $`h_{50}`$ is the Hubble constant in units of 50 km s<sup>-1</sup>Mpc<sup>-1</sup>. For a Hubble constant equal to 65 km s<sup>-1</sup>Mpc<sup>-1</sup>, $`\mathrm{\Omega }_b`$ = 0.040$`{}_{0.007}{}^{}{}_{}{}^{+0.009}`$. Our derived baryonic mass fraction is consistent with the one obtained by analysis of the Ly$`\alpha `$ forest in a cold dark matter cosmology. Depending on the intensity of diffuse UV radiation, the inferred lower limit is $`\mathrm{\Omega }`$$`h_{50}^2`$ = 0.05 โ 0.10 ( Weinberg et al. 1997; Bi & Davidsen 1997; Rauch et al. 1997), while Zhang et al. (1998) have derived 0.03 $``$$`\mathrm{\Omega }`$$`h_{50}^2`$ $``$ 0.08.
Finally, for the most consistent set of primordial abundances โ D from Levshakov et al. (1999), our above value for <sup>4</sup>He, and <sup>7</sup>Li from Vauclair & Charbonnel (1998) โ we derive an equivalent number of light neutrino species $`N_\nu `$ = 3.0$`\pm `$0.3 (2$`\sigma `$) (Izotov et al. 1999).
#### Acknowledgments.
We thank the partial financial support of NSF grant AST-96-16863 and an IAU Travel grant. We acknowledge useful conversations with M. Peimbert and B. Pagel and thank the organizers for a stimulating meeting in a superb locale.
## References
Bi, H., & Davidsen, A. F. 1997, ApJ, 479, 523
Bonifacio, P. & Molaro, P. 1997, MNRAS, 285, 847
Brocklehurst, M. 1972, MNRAS, 157, 211
Burles, S., & Tytler, D. 1998a, ApJ, 507, 732 1998, Phys. Rev. D, 58, 063506
Izotov, Y. I., & Thuan, T. X. 1998a, ApJ, 500, 188 (IT98a)
โโโ. 1998b, ApJ, 497, 227 (IT98b)
โโโ. 1999, ApJ, 511, 639
Izotov, Y. I., Thuan, T. X., & Lipovetsky, V. A. 1994, ApJ, 435, 647 (ITL94)
โโโ. 1997, ApJS, 108, 1 (ITL97)
Izotov, Y. I., Lipovetsky, V.A., Guseva, N.G., Kniazev, A. Y., Neizvestny, S. I. & Stepanian, J. A. 1993, Astron. Astrophys. Trans., 3, 193
Izotov, Y. I., Chaffee, F. H., Foltz, C. B., Green, R. F., Guseva, N. G., Thuan, T. X. 1999, ApJ, 527, 757
Kingdon, J., & Ferland, G. J. 1995, ApJ, 442, 714
Levshakov, S. A., Kegel, W. H., & Takahara, F. 1998, MNRAS, 302, 707
Markarian, B. E., Lipovetsky, V.A., Stepanian, J.A., Erastova, L. K., & Shapovalova, A. I. 1989, Commun. Special Astrophys. Obs., 62, 5
Olive, K. A., Skillman, E. D., & Steigman, G. 1997, ApJ, 483, 788 (OSS97)
Pagel, B. E. J., Simonson, E. A., Terlevich, R. J., & Edmunds, M. G. 1992, MNRAS, 255, 325
Pagel, B. E. J., Terlevich, R. J., & Melnick, J. 1986, PASP, 98, 1005
Peimbert, M., & Torres-Peimbert, S. 1974, ApJ, 193, 327
โโโ. 1976, ApJ, 203, 581
Peimbert, M. & Peimbert, A. 2000, this volume
Pinsonneault, M. H., Walker, T. P., Steigman, G., & Naranyanan, V. K. 1999, ApJ, 527, 180
Rauch, M., Miralda-Escudรฉ, J., Sargent, W. L. W., Barlow, T. A., Weinberg, D. H., Hernquist, L., Katz, N., Cen, R., & Ostriker, J. P. 1997, ApJ, 489, 7
Robbins, R. R. 1968, ApJ, 151, 511
Rood, R. T., Bania, T. M., Balser, D. S., & Wilson, T. L. 1998, Space Sci. Rev., 84, 185
Salzer, J. J., MacAlpine, G. M., & Boroson, T. A. 1989, ApJS, 70, 447
Smits, D. P. 1996, MNRAS, 278, 683
Songaila, A., Wampler, E. J., & Cowie, L. L. 1997, Nature, 385, 137
Stasiลska, G., & Schaerer, D. 1999, A&A, 351, 72
Vauclair, S., & Charbonnel, C. 1998, ApJ, 502, 372
Vรญlchez, J. M., & Pagel, B. E. J. 1988, MNRAS, 231, 257
Weinberg, D. H., Miralda-Escudรฉ, J., Hernquist, L., & Katz, N. 1997, ApJ, 490, 564
Zhang, Y., Meiksin, A., Anninos, P., & Norman, M. L. 1998, ApJ, 495, 63 |
warning/0003/hep-ex0003022.html | ar5iv | text | # Results from the Palo Verde Neutrino Oscillation Experiment
## I Introduction
Results of a long baseline study of $`\overline{\nu }_\mathrm{e}`$ oscillations at the Palo Verde Nuclear Generating Station are reported here. The work was motivated by the observation of an anomalous atmospheric neutrino ratio $`\nu _\mu /\nu _\mathrm{e}`$ reported in several independent experiments that can be interpreted as $`\nu _\mu `$$`\nu _\mathrm{e}`$ oscillations requiring large mixing. The mass parameter suggested by this anomaly is in the range of $`10^2<\mathrm{\Delta }m^2<10^3`$ eV<sup>2</sup> for two flavor neutrino oscillations.
The quantity $`\mathrm{\Delta }m^2`$, defined as the difference between the square of the masses of the mass eigenstates, and the mixing parameter $`\theta `$ are related to the transition probability $`P`$ for two-flavor $`\nu _a\nu _b`$ oscillations (see, for example, ) by:
$$P_{\mathrm{osc}}(\nu _a\nu _b)=\mathrm{sin}^22\theta \mathrm{sin}^2\left(\frac{1.27\mathrm{\Delta }m^2L}{E_\nu }\right),$$
(1)
where $`E_\nu `$ (MeV) is the neutrino energy, $`L`$ (m) is the sourceโdetector distance, and $`\mathrm{\Delta }m^2`$ is measured in eV<sup>2</sup>.
Exploring $`\mathrm{\Delta }m^2`$ down to 10<sup>-3</sup> eV<sup>2</sup> requires that the quantity $`L/E_\nu `$ (m/MeV) has a value of around 200. For reactor neutrinos ($`E_\nu `$5 MeV), a baseline of $`L`$1 km is adequate. Reactor experiments are generally well suited to study $`\overline{\nu }_\mathrm{e}`$ oscillations at small $`\mathrm{\Delta }m^2`$; however, they are restricted to the disappearance channel $`\overline{\nu }_\mathrm{e}\overline{\nu }_x`$.
Reactor antineutrinos have been used for oscillation studies with ever increasing $`\mathrm{\Delta }m^2`$ sensitivity since 1981. All of the experiments are based on the large cross section inverse beta decay reaction, $`\overline{\nu }_\mathrm{e}`$p$``$e<sup>+</sup>n. The correlated signature, a positron followed by a neutron capture, allows significant suppression of backgrounds. As the reactor $`\overline{\nu }_\mathrm{e}`$ yield and spectra are well known, a โnear detectorโ is not required. It is, however, important to control well the detector efficiency and backgrounds.
The mentioned considerations have led to the design of the Palo Verde and Chooz experiments, which have similar $`\mathrm{\Delta }m^2`$ sensitivities. While both experiments have pursued their goal of exploring the unknown region of small $`\mathrm{\Delta }m^2`$, recent data from Super-Kamiokande favor the $`\nu _\mu \nu _x`$ oscillation channel over $`\nu _\mu \nu _\mathrm{e}`$. This paper reports in greater detail results presented earlier and describes the detector calibration, background subtraction, and data analysis techniques used to extract results on neutrino oscillations.
## II The Experiment
### A The detector
The Palo Verde Nuclear Generating Station in Arizona, the largest nuclear power plant in the USA, consists of three identical pressurized water reactors with a total thermal power of 11.63 GW. The detector is located at a distance of 890 m from two of the reactors and 750 m from the third at a shallow underground site. The 32 meter-water-equivalent overburden entirely eliminates any hadronic component of cosmic radiation while reducing the cosmic muon flux to 22 $`\mathrm{m}^2\mathrm{s}^1`$. In order to reduce the ambient $`\gamma `$-ray flux in the laboratory all materials in and surrounding the detector were selected for low activity. The laboratory walls were built with an aggregate of crushed marble, selected for its low content of natural radioisotopes. Concentrations of 170, 750, and 560 ppb for <sup>40</sup>K, <sup>232</sup>Th, and <sup>238</sup>U were measured in the concrete resulting in a tenfold reduction of $`\gamma `$-ray flux when compared with locally available aggregate. A low <sup>222</sup>Rn concentration of about 20 Bq/m<sup>3</sup> in the lab air was maintained with forced ventilation. Temperature and humidity were controlled to ensure stable detector operation.
The segmented detector, shown in Fig. 1, consists of a 6$`\times `$11 array of acrylic cells dimensioned at 900 cm$`\times `$12.7 cm$`\times `$25.4 cm and filled with a total of 11.34 tons of liquid scintillator. A 0.8 m long oil buffer at the ends of each cell shields the central detector from radioactivity originating in the photomultiplier tubes (PMTs) and laboratory walls. The cells were made by cutting and bonding large 0.62 cm thick acrylic sheets. The total acrylic mass in the detector is 3.48 tons. Each cell is individually wrapped in 0.13 mm thick Cu foil to ensure light-tightness and is viewed by two 5-inch low activity PMTs, one at each end, housed in mu-metal boxes. The target cells are suspended on rollers held in place by thin sheet metal hangers. All structural materials were dimensioned as lightly as possible to minimize dead material between cells. Each cell can be individually removed from the mechanical structure for maintenance. The detector is oriented such that the $`\overline{\nu }_\mathrm{e}`$ flux is perpendicular to the long axis of the cells.
The liquid scintillator is composed of 36% pseudocumene, 60% mineral oil, and 4% alcohol, and is loaded with 0.1% Gd by weight. This formulation was chosen to yield long light transmission length ($`11.5\pm 0.1`$ m at 440 nm), good stability, high light output, and long term compatibility with acrylic. Details of the scintillator development have been published elsewhere .
The central volume is surrounded on the sides by a 1 m buffer of high purity deionized water (about 105 tons) contained in steel tanks which, together with the oil buffers at the ends of the cells, serve to attenuate gamma radiation from the laboratory walls as well as neutrons produced by cosmic muons passing outside of the detector. The low Z of water minimizes the neutron production by nuclear capture of stopped muons inside the detector and has a high efficiency for neutron thermalization.
The outermost layer of the detector is an active muon veto counter, providing 4$`\pi `$ coverage. It consists of 32 twelve meter-long PVC tanks (from the MACRO experiment) surrounding the detector longitudinally, and two endcaps. The endcaps are mounted on a rail system to allow access to the central detector. The horizontal tanks are read out by two 5-inch PMTs at each end; the vertical tanks are equipped with one 8-inch PMT at each end while the endcaps use 3-inch PMTs. The liquid scintillator used in the veto is a mixture of 2% pseudocumene and 98% mineral oil, with a light attenuation length at 440 nm in excess of 12 m.
A schematic of the central detectorโs front-end electronics is shown in Fig. 2. Each channel can be digitized by either of two identical banks of electronics. The dual bank system allows both parts of the sequential inverse beta decay event to be recorded with no deadtime by switching between banks. Due to the large dynamic range of energy in the data of interest (40 keV to 10 MeV, or 1 to 250 photoelectrons typically), each PMT has both a dynode and anode output connected to ADCs, as well as three discriminator thresholds for the trigger and TDCs. The higher TDC threshold serves to avoid crosstalk from large signals in adjacent channels while the lower threshold allows timing information to still be available at the single photoelectron level. The relative time of arrival from each end of a cell is used to reconstruct longitudinal position. The measured PMT pulse charge at each end, corrected for light attenuation based on the distance traveled in the cell, allows energy reconstruction.
Each cell is connected to the trigger via the or of the discriminated signals from the two PMTs. Signals are tagged according to two thresholds: a high threshold corresponding to $``$600 keV for energy deposits in the middle of the cell and a low threshold corresponding to $``$40 keV, or one photoelectron at the PMT. The low trigger threshold also serves as the lower TDC threshold. The trigger, which has a decision time for each event of around 40 ns, uses a Field Programmable Gate Array to search for patterns of energy deposits in the central detector, and can be reprogrammed easily to change trigger conditions as needed for calibrations.
A veto signal disables the central detector trigger for 10 $`\mu `$s following the passage of a muon to avoid most related activity. Typical veto rates are $``$2 kHz. With each event, the time and hit pattern of the previous muon in the veto counter is recorded along with information as to whether or not the muon passed through the target cells. The veto inefficiency was measured to be (4$`\pm `$1)% for stopping muons (one hit missed) and (0.07$`\pm `$0.02)% for through-going muons (two hits missed). We note that the small size of this second quantity with respect to the first is due to correlations between incoming and outgoing muons as confirmed by a simple Monte Carlo model.
### B The $`\overline{\nu }_\mathrm{e}`$ signal
The $`\overline{\nu }_\mathrm{e}`$ signal is detected via the reaction $`\overline{\nu }_\mathrm{e}`$p$``$ne<sup>+</sup> as illustrated in Fig. 17 further below along with the dominant backgrounds. Signal events consist of a pair of time-correlated subevents: (1) the positron kinetic energy ionization and two annihilation $`\gamma `$โs forming the prompt part and (2) the subsequent capture of the thermalized neutron on Gd forming the delayed part. By loading the scintillator with 0.1% Gd, which has a high thermal neutron capture cross section, the neutron capture time is reduced to $``$27 $`\mu `$s from $``$170 $`\mu `$s for the unloaded scintillator. Furthermore, Gd de-excites by releasing an 8 MeV $`\gamma `$ cascade, whose summed energy gives a robust event tag well above natural radioactivity. In contrast, neutron capture on protons releases only a single 2.2 MeV $`\gamma `$.
Background is rejected at trigger level using the detector segmentation by looking for coincidences of energy deposits matching the pattern of inverse beta decay. Each of the subevents of a $`\overline{\nu }_\mathrm{e}`$ signal is triggered by scanning the detector for a pattern of three simultaneous hits in any 3$`\times `$5 subset of the cell array. This threefold coincidence, called a triple, must consist of at least one high trigger hit, due to either the positron ionization or neutron capture cascade core, and at least two additional low trigger hits, resulting from either positron annihilation $`\gamma `$โs or neutron capture shower tails. The use of identical trigger requirements for the two triples is found to give rise to close to an optimal signal to noise ratio. Five $`\mu `$s after finding an initial triple, the trigger begins searching for a delayed triple. The blank time suppresses possible false signals from PMT afterpulsing. If two triples are found within 450 $`\mu `$s of each other, the candidate $`\overline{\nu }_\mathrm{e}`$ event is digitized for offline analysis.
### C Expected $`\overline{\nu }_\mathrm{e}`$ interaction rate
In order to calculate the expected $`\overline{\nu }_\mathrm{e}`$ interaction rate in the detector, the status of the three reactors is tracked daily, and the fission rates in the cores are calculated based on a simulation code provided by the manufacturer of the reactors. This code uses as input the power level of the reactors, various parameters measured in the primary cooling loop, and the original composition of the core fuel elements.
The output of the core simulation has been checked by measuring isotopic abundances in expended fuel elements in the core; errors in fuel exposure and isotopic abundances are estimated to cause $`<0.3`$% uncertainty in the $`\overline{\nu }_\mathrm{e}`$ flux estimate. Of the four isotopes โ <sup>239</sup>Pu, <sup>241</sup>Pu, <sup>235</sup>U, and <sup>238</sup>U โ whose fissions produce virtually all of the thermal power as well as neutrinos, measurements of the neutrino yield per fission and energy spectra exist for the first three. The <sup>238</sup>U yield, which contributes 11% to the final $`\overline{\nu }_\mathrm{e}`$ rate, is calculated from theory. When the same theoretical method was used to calculate the spectrum from the other three isotopes, the theory agreed with experimental results within 10%. The contribution of <sup>238</sup>U fission to the overall uncertainty in $`\overline{\nu }_\mathrm{e}`$ rate is therefore expected to be $``$1%.
This calculated $`\overline{\nu }_\mathrm{e}`$ flux is then used to compute the expected rate of $`\overline{\nu }_\mathrm{e}`$ candidates $`N_{\overline{\nu }_\mathrm{e}}`$ at the detector as a function of the oscillation parameters $`\mathrm{\Delta }m^2`$ and $`\mathrm{sin}^22\theta `$:
$`N_{\overline{\nu }_\mathrm{e}}=n_\mathrm{p}{\displaystyle }dE_{\overline{\nu }_\mathrm{e}}\sigma (E_{\overline{\nu }_\mathrm{e}})\eta (E_{\overline{\nu }_\mathrm{e}})\times `$ (3)
$`{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{_{\overline{\nu }_\mathrm{e},i}(E_{\overline{\nu }_\mathrm{e}})\left(1P_{\mathrm{osc},i}(\mathrm{\Delta }\mathrm{m}^2,\mathrm{sin}^22\theta ,L_i,E_{\overline{\nu }_\mathrm{e}})\right)}{4\pi L_i^2}}`$
where $`\sigma (E_{\overline{\nu }_\mathrm{e}})`$ is the inverse beta decay cross section, $`\eta (E_{\overline{\nu }_\mathrm{e}})`$ is the (energy dependent) detector efficiency, $`n_\mathrm{p}`$ is the number of target free protons, and $`_{\overline{\nu }_\mathrm{e},i}`$ is the source strength of reactor $`i`$ at distance $`L_i`$ with oscillation probability $`P_{\mathrm{osc},i}`$. In Fig. 3 we show the energy spectrum of the $`\overline{\nu }_\mathrm{e}`$โs emitted by a reactor, the $`\overline{\nu }_\mathrm{e}`$ (energy) differential cross section in the detector and the actual interaction rate in the detector target before detector efficiency corrections, referred to here as $`R_{\overline{\nu }_\mathrm{e}}`$ (obtained by setting $`\eta (E_{\overline{\nu }_\mathrm{e}})`$=1). The energy spectrum actually measured in the detector is the energy of the positron created by the inverse beta decay. This spectrum is approximately $`E_{\overline{\nu }_\mathrm{e}}1.8`$ MeV, slightly modified by the kinetic energy carried away by the neutron ($``$50 keV).
Previous short baseline experiments which measured the rate of $`\overline{\nu }_\mathrm{e}`$ emission by reactors have found good agreement between calculated and observed neutrino flux by using largely the same method of calculation. A high statistics measurement at Bugey, in particular, found excellent agreement both in spectral shape ($`\chi ^2/\mathrm{n}.\mathrm{d}.\mathrm{f}.`$=9.23/11) and in absolute neutrino yield (agreement better than 3%, dominated by systematic errors). These previous generation experiments prove that the reactor antineutrino spectrum, i.e. the $`\overline{\nu }_\mathrm{e}`$ flux at the distance $`L`$=0, is well understood.
The expected $`\overline{\nu }_\mathrm{e}`$ interaction rate in the whole target, both scintillator and the acrylic cells, is plotted in Fig. 4 for the case of no oscillations from July 1998 to October 1999. Around 220 interactions per day are expected with all three units at full power. The periods of sharply reduced rate occurred when one of the three reactors was off for refueling, the more distant reactors each contributing approximately 30% of the rate and the closer reactor the remaining 40%. The short spikes of decreased rate are due to short reactor outages, usually less than a day. The gradual decline in rate between refuelings is caused by fuel burnup, which changes the fuel composition in the core and the relative fission rates of the isotopes, thereby affecting slightly the spectral shape of the emitted $`\overline{\nu }_\mathrm{e}`$ flux.
## III Calibration
In order to maintain constant data quality during running, a program of continuous calibration and monitoring of all central detector cells is followed. Blue LEDs installed inside each cell are used for relative timing and position calibration. Optical fibers at the end of each cell, also illuminated by blue LEDs, provide information about PMT linearity and short term gain changes. LED and fiberoptic scans are performed once a week. Radioactive sources are used to map the light attenuation in each cell, for absolute energy calibration, and to determine detection efficiencies for positrons and neutrons. A complete source scan is undertaken every 2โ3 months.
### A LED and optical fiber calibrations
As seen in Fig. 1, every cell of the central detector has two LEDs, one at each end at a distance of 90 cm from the PMTs. These blue LEDs, which provide fast light pulses with a rise time comparable to scintillation light, are used for timing calibrations needed for position reconstruction along the cellโs axis.
The difference in pulse arrival time between the two PMTs of a cell $`\mathrm{\Delta }t`$ is described as a function of the position $`z`$ with an effective speed of light $`c_{\mathrm{eff}}`$, an offset $`z_0`$ and a small nonlinear correction $`f(1/Q)`$:
$$\mathrm{\Delta }t=(zz_0)/c_{\mathrm{eff}}+f(1/Q_n,1/Q_f).$$
(4)
The correction $`f(1/Q_n,1/Q_f)`$, a function of both near and far PMT pulse charge $`Q`$, describes the dependence of the pulse height due to time-walk in the leading edge discriminators used in the front-end electronics. To extract these calibration parameters and compensate for the time-walk effect, a third order polynomial is fit to $`\mathrm{\Delta }t`$ versus $`1/Q`$ (see Fig. 5). The intercepts at $`1/Q=0`$ for the two LED positions provide $`c_{\mathrm{eff}}`$ and $`z_0`$, while the slopes are used to parameterize the time-walk correction.
In order to check the suitability of longer wavelength 470 nm LED light to measure timing properties of $``$425 nm scintillation light, data taken with a <sup>228</sup>Th source at several longitudinal positions were reconstructed with the LED timing calibration parameters. Comparing the reconstructed positions with the actual source positions, the effective speed of light measured with the LED system was found to be on average 3.6% lower than that with the sources. A simulation of the light transport in a cell with various indices of refraction and attenuation lengths of the scintillator suggested that the small discrepancy in $`\mathrm{\Delta }t`$ between LED and scintillation light was due to the difference in attenuation length. The correction factor was found to be constant over several months. Weekly LED scans are therefore used to correct for short term variations in $`\mathrm{\Delta }t`$ and a constant correction factor is applied to the effective speed of light.
The fiberoptic system includes 15 blue LEDs, each illuminating a bundle of 12 fibers. The light output of each LED is measured in two independent reference cells with PMTs checked to be linear over the whole dynamic range of the LEDs. By taking a run which scans through all light intensities and mapping each PMTโs response relative to the reference cells, the nonlinear energy response of the PMTs is calibrated. Low intensities are used to determine the single photoelectron gain of each PMT, which is used to correct for changes from the nominal gain setting of $`4\times 10^7`$.
### B Scintillator transparency and energy scale calibration
In addition to weekly LED and fiberoptic calibrations, the energy response of the scintillator is measured every three months using a set of sealed radioactive sources. Eighteen 2.4 mm diameter tubes run along the length of the detector allowing insertion of the sources adjacent to any cell at any longitudinal position. The response of each PMT as a function of longitudinal position is measured by recording the Compton spectrum from the 2.614 MeV $`\gamma `$ of a <sup>228</sup>Th source at seven different locations along each cell.
Monte Carlo simulation found that the half maximum of a Gaussian function fitted to the Compton spectrum is relatively independent of resolution; this point is therefore used as the benchmark of the cell response. The response versus distance from the PMT, shown in Fig. 6 for one cell, is then fit to the phenomenological function $`\mathrm{exp}(p_0+p_1z)+\mathrm{exp}(p_2+p_3z)/z`$, where $`z`$ is source longitudinal distance from the PMT. The effective attenuation length of the scintillator (including multiple total reflection on the acrylic walls) is generally between 3โ4 m and over a year was found to change on average $``$1 mm/day, demonstrating that the Gd scintillator was remarkably stable.
The overall energy scale was determined from the position of the 1.275 MeV peak of a <sup>22</sup>Na source, and then verified by taking data with several $`\gamma `$ sources in different energy ranges: <sup>137</sup>Cs (0.662 MeV), <sup>65</sup>Zn (1.351 MeV), <sup>228</sup>Th (2.614 MeV), and the capture of neutrons (8 MeV) from an Am-Be source. The gamma cascade from neutron capture was modeled according to measurements of the emitted spectrum. In contrast to homogeneous detectors which measure total absorption energy peaks, 25% of the detector target mass consists of the inert acrylic of the cell walls, which absorbs some energy. The Monte Carlo simulation was therefore used to find the correct final distributions of energy detected from single and multiple scattering of the $`\gamma `$โs. The total energy reconstructed for data and Monte Carlo for each source is plotted in Fig. 7. The data were matched with Monte Carlo simulation for the <sup>22</sup>Na spectrum in Fig. 12 to find the overall energy scale and to the spectra in Fig. 7 to assure that the scintillator response is linear over the energies of interest. The light yield after PMT quantum efficiency was found to be $``$50 photoelectrons per MeV in the center of the cells. The agreement for three of the four sources in Fig. 7 is good, the exception being <sup>228</sup>Th, in which the data has a consistently higher Compton scattering peak than Monte Carlo predicts. This discrepancy is consistent across all the data taken and therefore does not affect the scintillator transparency calibration.
### C Monte Carlo simulation
The $`\overline{\nu }_\mathrm{e}`$ efficiency of the detector is a relatively strong function of event location in the detector and, to a lesser extent, of time due to scintillator aging. A further complication comes from the trigger efficiency being a function of threshold (voltage) while only energy (charge) is measured. For this reason a Monte Carlo model which included a detailed simulation of the detector response, including the PMT pulse shape, is used for an estimate of the overall efficiency for $`\overline{\nu }_\mathrm{e}`$ detection. A variety of measurements was performed to crosscheck that the Monte Carlo accurately models the detector response.
The physics simulation program is based on geant 3.21. This code contains the whole detector geometry and simulated the energy, time, and position of energy deposits in the detector. Hadronic interactions are simulated by gfluka and the low energy neutron transport by gcalor. Scintillator light quenching, parameterized as a function of ionization density, is included in the simulation.
The event reconstruction program reads the output of this physics simulation and then applies the second step of the Monte Carlo, the simulation of the detector response as PMT pulses which are then converted into time and amplitude digitizations and trigger hits. A logical scheme of this detailed detector simulation is shown in Fig. 8.
The calibrations discussed above empirically provide the scintillator light yield (photoelectron/MeV) and attenuation function for each cell, which in turn provide the number of photoelectrons $`\overline{N}_{p.e.}`$ expected for a simulated energy deposit. The total charge of the pulse then follows from sampling a Poisson distribution with mean $`\overline{N}_{p.e.}`$ and folding the number of simulated photoelectrons $`N_{p.e.}`$ with the PMTโs nominal gain, first stage gain variance (10$`N_{p.e.}`$), and cell-to-cell energy scale calibration uncertainty (10%).
To simulate the pulse shape, an arrival time is assigned to each photoelectron, and individual photoelectron pulses (whose shape is derived from real data) are summed into a final pulse. The calculated arrival time of each photon is a combination of two processes, scintillator de-excitation and propagation along the cell. The latter distribution is parameterized by the distance traveled to the PMT, larger distances giving larger variances, using a light transport simulation of $`2\times 10^7`$ photons. The resulting pulse is then analyzed to extract TDC and trigger hits.
The Monte Carlo threshold simulation, position reconstruction, and positron and neutron efficiency predictions were checked using calibration data. The trigger threshold simulation for each cell was compared to data taken with a <sup>22</sup>Na $`\beta ^+`$ source near the center of each cell. The trigger conditions were loosened for these data, a single low hit producing a trigger and the event tagged if a high threshold was crossed. By plotting the reconstructed energy for each event versus the efficiency for a high trigger tag, an effective high trigger threshold in MeV for that location in the cell was determined. The low threshold was measured similarly. The Monte Carlo pulse shape parameters were tuned to these data. A typical cellโs trigger threshold efficiency as a function of energy is shown in Fig. 9 for both data and the Monte Carlo. The trigger threshold, defined as the energy at 50% efficiency, is also plotted for all 66 cells. On average, the thresholds agree to within 1%.
TDC thresholds were checked by the same algorithm, plotting the threshold hit efficiency versus reconstructed energy. A more direct check of the TDC simulation, however, compares the position reconstruction for data and Monte Carlo simulation. Fig. 10 shows the longitudinal position of the third largest energy deposit in each event for a <sup>22</sup>Na calibration run, representing the position reconstruction of the energy deposited by one of the two positron annihilation $`\gamma `$โs. Since these energy deposits tend to be small ($``$100 keV), some fraction of them have one or both PMTโs responses below the TDC low threshold. These events constitute the tails of the distribution in Fig. 10 since only the relative signal amplitude was used for position reconstruction. The narrower central peak is populated by events with TDC information available. The simulation and data agree well, in both resolution and relative frequency of the two cases.
### D $`\overline{\nu }_\mathrm{e}`$ detection efficiency
The absolute efficiency of the detector for positron annihilations and neutron captures was verified using <sup>22</sup>Na and Am-Be sources respectively. The <sup>22</sup>Na source emits a 1.275 MeV primary $`\gamma `$ which is accompanied 90% of the time by a low energy positron which annihilates in the source capsule. The primary $`\gamma `$ can mimic the positron ionization of a low energy $`\overline{\nu }_\mathrm{e}`$ event. This deposit, in conjunction with the positronโs annihilation $`\gamma `$โs, closely approximates the positron portion of a $`\overline{\nu }_\mathrm{e}`$ event near the trigger threshold.
In two rounds of data taking, 10 months apart, the <sup>22</sup>Na source was inserted into the central detector at 35 locations chosen to provide a sampling of various distances from the PMTs and edges of the fiducial volume. The source activity is known to 1.5%, allowing determination of an absolute efficiency. After applying the offline selections used for $`\overline{\nu }_\mathrm{e}`$ prompt triples and correcting for detector DAQ deadtime, the measured absolute efficiency was compared with the Monte Carlo prediction; the results are summarized in the top portion of Fig. 11. Good agreement is seen in the average efficiency over all runs, and run by run agreement was 11%.
The energy spectra predicted by the simulation and measured in the data for the <sup>22</sup>Na runs were compared. The total energy seen in all cells and the energy detected in the three most energetic hits are plotted in Fig. 12. The trigger thresholds can be seen in the spectra: the high trigger threshold is the rising edge at around 0.5 MeV in the spectrum of the most energetic hit (E<sub>1</sub>), and the low trigger threshold is the rising edge at around 50 keV of the third most energetic hit (E<sub>3</sub>).
A similar procedure was used to check the neutron capture detection efficiency. The Am-Be neutron source is attached to one end of a thin (7.5mm) NaI(Tl)-detector, which tagged the 4.4 MeV $`\gamma `$ emitted in coincidence with a neutron. The NaI(Tl) tag forces the digitization of the 4.4 MeV $`\gamma `$ as the prompt part of an event and opens a 450 $`\mu `$s window for neutron capture; this is the same coincidence window used in the $`\overline{\nu }_\mathrm{e}`$ runs.
All neutron cuts used for the $`\overline{\nu }_\mathrm{e}`$ data selection were applied, and the resulting detection efficiency was corrected for detector deadtime and a small random coincidence background. On average, the Monte Carlo efficiency predictions agrees well over the 25 locations tested with a run by run agreement of better than 4%, as shown in the bottom of Fig. 11.
As with the <sup>22</sup>Na runs, the energy spectra predicted by the simulation and measured in the data were compared. The total energy seen in all cells and the energy detected in the three most energetic hits is plotted in Fig. 13. Note the small peak in $`E_{\mathrm{total}}`$ at $``$2 MeV arising from neutrons being captured on hydrogen. The differences in data versus Monte Carlo spectra for <sup>22</sup>Na and Am-Be were taken into account in estimating systematic errors.
The Am-Be source emits neutrons with kinetic energies up to 10 MeV, creating proton recoils in the detector scintillator in coincidence with the NaI(Tl) induced trigger. By digitizing any energy deposits seen during the neutron release, the high ionization density of these recoiling protons was used in setting the parameters which control scintillator light quenching in the simulation.
The above crosschecks verify our ability to accurately generate the events, model the detector response, reconstruct the events, and correctly calculate the livetime of the data acquisition (DAQ) system. Taken together these procedures complete the task of estimating our $`\overline{\nu }_\mathrm{e}`$ efficiency.
The Monte Carlo simulation for $`\overline{\nu }_\mathrm{e}`$ events models the expected interactions throughout the entire target, including the acrylic walls of the cells, since there is significant efficiency for inverse beta decay originating in the acrylic. The Monte Carlo simulation yields an average efficiency over the entire detector as a function of $`\overline{\nu }_\mathrm{e}`$ energy. The efficiency from the simulation is folded with the incident $`\overline{\nu }_\mathrm{e}`$ spectrum (which may be distorted by oscillations depending on the hypothesis tested), to get the effective efficiency.
### E An independent reconstruction and Monte Carlo
A parallel and independent event reconstruction and the simulation of the detector response has been developed. This second version follows the same general outline of detector calibration, event reconstruction, and simulation described above, but differs in the algorithms and parameterizations used. Major differences include:
* The functional form for the scintillator light attenuation is the sum of two exponentials $`\mathrm{exp}(p_0+p_1z)+\mathrm{exp}(p_2+p_3z)`$, without $`z^1`$ in the second term.
* The cell response benchmark is the 70% maximum rather than half maximum of the fitted Compton scattering spectrum.
* A different parameterization is used for the linearity correction of the dynode signals.
* The low threshold parameters are tuned to the TDC hit efficiencies rather than trigger efficiencies as discussed above.
* An alternate algorithm for simulating the PMT pulse shape was developed and tuned to observed PMT pulse characteristics.
These differences manifest themselves as slightly different $`\overline{\nu }_\mathrm{e}`$ efficiency predictions and $`\overline{\nu }_\mathrm{e}`$ candidate rates in the data. Tests with radioactive sources have been performed to evaluate the quality of the second data reconstruction. The <sup>22</sup>Na and Am-Be efficiency runs shown in Fig. 11 were reconstructed by the second analysis to test its efficiency prediction throughout the detector. The ratio of predicted to observed efficiencies over all the e<sup>+</sup> and neutron runs for the first reconstruction (1) and the second reconstruction (2) are plotted in Fig. 14. While the results presented in the analyses below come from the first reconstruction code described above, the development of a second simulation and event reconstruction offers a useful crosscheck of the systematic uncertainties of the results. The differences between the two analyses were used to corroborate the estimate of systematic errors.
## IV $`\overline{\nu }_\mathrm{e}`$ selections and backgrounds
### A $`\overline{\nu }_\mathrm{e}`$ selection
The trigger rate for time-correlated events (two triples occurring within 450 $`\mu `$s) is $``$1 Hz. Most of those events are random coincidences of two uncorrelated triple hits, which occur individually at a rate of $``$50 Hz, mostly from natural radioactivity. In order to select neutrino events, the following offline cuts are applied:
* The energy reconstructed in both prompt and delayed triples has at least one hit with $`E>`$1 MeV and at least two additional hits with $`E>`$30 keV. No single hit was allowed to be greater than 8 MeV.
* The prompt triple is required to resemble a positron, i.e. annihilation $`\gamma `$โs each less than 600 keV, and together less than 1.2 MeV. (This cut is the only one which treats the two triples asymmetrically).
* At least one of the two triples in the event has more than 3.5 MeV of reconstructed energy for rejection of $`\gamma `$ backgrounds.
* The prompt and delayed portions of the event are correlated in space and time (within 3 columns, 2 rows, one meter longitudinally, and 200 $`\mu `$s).
* The event started at least 150 $`\mu `$s ($``$5 neutron capture times) after the previous veto tagged muon activity.
The trigger and selection efficiencies are summarized in the first two columns of Table I.
In addition to corrections for selection cut efficiency and trigger efficiency, detector livetime is a substantial correction to the number of neutrinos seen and deserves some comment. Deadtime comes from two sources, the DAQ and the muon veto. DAQ livetime is the ratio of the number of triples the DAQ was available for digitizing to the total number of triples the trigger saw. These numbers are available from trigger scalers. The trigger livetime was measured to be $`>`$99.9%. The DAQ live time varies with the triple rate, and for the four data periods was determined to be 73.2%, 74.4%, 92.3%, and 91.8% for 1998 full power, 1998 refueling, 1999 full power, and 1999 refueling, respectively. The higher livetime in 1999 is due to improvements made in the trigger conditions.
The muon deadtime can be further divided into two contributions: 150 $`\mu `$s of deadtime caused by each muon, which at $`R_\mu =`$ 1990 Hz left the detector live 74.2% of the time; and muons which interrupted a neutrino event between the positron and the neutron capture, which estimated from the fit parameters of the Monte Carlo capture time left 92.5% of events uninterrupted. The total uncertainty in the calculation of detector deadtime is less than 1%
### B Backgrounds
Backgrounds can be separated into two types: correlated and uncorrelated. Uncorrelated background events are due to unrelated triple hits which randomly coincided in the time window allowed. Although most of the events collected were random coincidences, almost all of this type of background is removed by requiring at least one subevent to have more than 3.5 MeV of reconstructed energy. These events do not have a time correlation between prompt and delayed subevents (inter-event time) characteristic of neutron capture. They have instead a longer time correlation determined by the probability that the veto detected no muon between the prompt and delayed random triples. At a 2 kHz muon rate, this background is seen as a 500 $`\mu `$s tail under the normal neutron capture distribution. By looking at the inter-event times of the candidate $`\overline{\nu }_\mathrm{e}`$ events at longer time scales, this background can be measured.
The inter-event time distribution after all neutrino selections (except the time correlation cut) is shown in Fig. 15. The Monte Carlo for a pure neutron capture sample is empirically fitted to the sum of two exponentials. There are two time constants due to the inhomogeneity of the target: neutrons which remain in the scintillator have a 27 $`\mu `$s capture time, whereas those which enter the acrylic have a longer capture time due to the absence of Gd. The data inter-event time distribution is fitted to a function of three exponentials with fixed time constants consisting of the Monte Carlo fit $`\tau `$โs multiplied by a third time constant of 500 $`\mu `$s. Integrating the resulting 500 $`\mu `$s exponential of the uncorrelated background in the signal region gives an estimate of 4.1$`\pm `$0.2 events per day, or 9% of the $`\overline{\nu }_\mathrm{e}`$ candidates being uncorrelated background events.
To measure the uncorrelated background in smaller parts of the data set, the statistical accuracy of the three exponential fit method becomes unacceptably poor. A simpler method is therefore used in conjunction with the above fit. For inter-event times longer than 200 $`\mu `$s, the $`\overline{\nu }_\mathrm{e}`$ candidates are dominated by uncorrelated backgrounds. The integrated number of candidates from 200โ400 $`\mu `$s is scaled to estimate the number underneath the signal region ($`<`$200 $`\mu `$s). Using the scaling from the fit of the entire data set shown in Fig. 15, the uncorrelated background was measured in approximately month-long intervals as shown in Fig. 16. For both the 1998 and 1999 data sets, the rates are found to be stable within statistical errors.
Correlated backgrounds have the neutron capture inter-event time structure of the $`\overline{\nu }_\mathrm{e}`$ candidates. These events come mainly from cosmic muon induced fast neutrons from spallation or muon capture, as shown schematically in Fig. 17. These fast neutrons can either (1) induce more neutrons via spallation, two of which can be captured in the detector with one capture mimicking a positron signature; or (2) they can cause proton recoil patterns in the central detector which appear as a positron signature and then get captured. Spallation neutrons originate from muons passing through the walls of the lab without hitting the veto detector or from muons passing through the detector shielding undetected by the veto. Muon capture neutrons mainly originate from muons stopping in the water buffer without registering in the veto.
To illustrate some properties of correlated background, Fig. 18 shows the time elapsed since the previous veto hit for $`\overline{\nu }_\mathrm{e}`$ candidates, with all selection cuts applied except that of the previous muon timing. This distribution is fit to a three exponential function analogous to that used for the inter-event time fits. The two time constants for neutron capture are not identical to those for $`\overline{\nu }_\mathrm{e}`$ events, but tend to be smaller since after passage of a muon there are often more than one neutron in the detector to be captured. The third exponential time constant is again constrained to 500 $`\mu `$s as expected in a random sampling of events unrelated to the previous muon. Since at very short times there are other contributions such as muon decay, times less than 15 $`\mu `$s are excluded from the fit. Muon-induced-neutron backgrounds dominate the candidates in the first 150 $`\mu `$s after the previous tagged muons, motivating the selection cut on $`\mu `$ timing.
In order to show that the correlated background was constant in time, the previous muon time cut was disabled and a plot was made of the $`\overline{\nu }_\mathrm{e}`$ candidate rate versus time, as shown in Fig. 19. When fit to a constant for each year, a $`\chi ^2`$/n.d.f. of 382.7/371 is obtained, which has a 33% likelihood, indicating that the detector efficiency for correlated backgrounds was stable during each yearโs data taking.
Aside from the detector efficiency for background, however, a loss of veto efficiency could also cause a fluctuation in background. (The rates fit in Fig. 19 are with the muon timing selection disabled, and hence do not vary with veto inefficiency.) To track veto efficiency, the veto hit patterns recorded with each event are used. If a muon hit was recorded only on the bottom of the veto, where only exiting muons are seen, then the muon must have entered the veto without recording a hit. By measuring the percentage of these events a one hit missed veto inefficiency of ($`4\pm 1`$)% is found as mentioned above. The through-going (two hits missed) veto inefficiency is measured to be $`0.07\pm 0.02\%`$ by looking at the rate of $`\mu `$ tracks triggered on in the central detector. These inefficiencies were tracked in time to assure their stability.
### C Neutronโ$`\overline{\nu }_\mathrm{e}`$ direction correlation
The neutrons produced in the inverse beta decays will have momenta slightly biased away from the source, whereas no correlation is expected for background. This effect is the consequence of momentum conservation which requires that the neutron should always be emitted in the forward hemisphere with respect to the incoming $`\overline{\nu }_\mathrm{e}`$. Such a correlation has been observed already in the Gรถsgen experiment and again at Chooz. The theoretical treatment of the effect can be found in .
The signal to background ratio can be independently verified using this effect. The $`\overline{\nu }_\mathrm{e}`$ source is to the left of the detector in Fig. 1. The relative horizontal location (relative column in the target cell array) of neutron capture cascade cores versus positron ionizations for data and the simulation of the $`\overline{\nu }_\mathrm{e}`$ signal are plotted in Fig. 20. Defining the asymmetry $`A_{\mathrm{data}}=\frac{RL}{R+L}`$ in terms of the number of neutrons captured one column away from the source $`R`$ and one column toward the source $`L`$, a slight asymmetry $`0.050\pm 0.017`$ is found in the data, at 2.9 $`\sigma `$ significance. Using the Monte Carlo simulation which gives $`A_{\mathrm{MC}}=0.134`$ to estimate the portion of the data consisting of $`\overline{\nu }_\mathrm{e}`$ signal and assuming the background to be symmetric in this variable, an effective signal to noise ratio
$$\frac{S}{N}=\frac{A_{\mathrm{data}}}{A_{\mathrm{MC}}A_{\mathrm{data}}}=0.6_{0.3}^{+0.4}$$
(5)
is found. This value agrees well with the ratio of $`0.81\pm 0.03`$ found with the swap analysis method described below.
## V Analysis
The data set presented here was taken from July 1998 to September 1999 in 373 short runs, each on average about 12 hours long. In 1998, 35.97 days of data were taken with the three reactors at full power and 31.35 days with one of the reactors at a distance of 890 m off for refueling. The detector was then taken offline in Jan/Feb of 1999, when DAQ improvements were made to increase livetime, and the high trigger thresholds were lowered by 30% to increase trigger efficiency. The 1999 data set includes 110.95 days with all three reactors at full power and 23.40 days with the 750 m baseline reactor off for refueling. Thus, the entire data set has four distinctly different periods, with three different baseline combinations and neutrino fluxes.
After all selection cuts there is still substantial background in the remaining data set. The correlated background, coming mainly from muon induced neutrons, is difficult to predict and subtract. The yield and spectrum of neutron spallation is a function of muon flux and energy, which in turn is a function of depth. While some measurements of fast neutron spectra and fluxes have been done in the past, there is no model which can consistently predict the fast neutron production. Below we present three methods used to extract the $`\overline{\nu }_\mathrm{e}`$ signal from data.
### A Analysis with the on-off method
The conceptually simplest method of subtracting background is to take advantage of periods of reduced power levels of the reactor source. Ideally all three reactors would be down at once allowing for a direct measurement of the background. However, in practice only one of the three Palo Verde reactors was refueled at any given time. These reduced power periods occurred twice annually for about a month. Each yearโs data set is treated independently, subtracting 1998 off from 1998 on and 1999 off from 1999 on, since the efficiency of the detector changed between the two years. By subtracting these data taken at reduced flux from the full flux data, a pure neutrino sample is retrieved albeit containing the statistical power of only a small portion of the potential data set: the subtraction is limited by the two months of refueling time and treats the $`\overline{\nu }_\mathrm{e}`$ flux from the two reactors still at full power as background.
The primary concern arising from use of this method, aside from the loss of statistics, is guaranteeing that the background rates during the on and off periods were stable. Both correlated and uncorrelated backgrounds were carefully tracked to ensure stability as discussed above.
The numerical results of this analysis of the total rate are summarized in Table II. After correcting for efficiency (for the no-oscillations scenario) and livetime, the data sets were subtracted to find observed neutrino interaction rates in the detector. No significant deviation from the expected neutrino interaction rates was found at either baseline distance.
The results from the alternate reconstruction (2) for this analysis are shown in Table III for comparison. This analysis selects about 5% more candidates, but also gives a correspondingly higher efficiency. For this analysis, the uncorrelated background was measured and removed from the data before the subtraction.
In order to test the results for oscillation hypotheses in the two flavor $`\mathrm{\Delta }m^2\mathrm{sin}^22\theta `$ plane, a $`\chi ^2`$ analysis is performed comparing the calculated $`R_{\mathrm{calc},ij}`$ and observed $`R_{\mathrm{obs},ij}`$ spectra divided into 1 MeV bins $`j`$ for each year $`i`$. The spectra used are the prompt energies of the two subtracted data sets. At each point in the oscillation parameter plane, taking into account the changes in detector efficiency due to distortions of the neutrino spectrum, the quantity
$$\chi ^2=\underset{i=1}{\overset{2}{}}\underset{j}{\overset{E_{\mathrm{bins}}}{}}\frac{\left(\alpha R_{\mathrm{calc},ij}R_{\mathrm{obs},ij}\right)^2}{\sigma _{ij}^2}+\frac{(\alpha 1)^2}{\sigma _{\mathrm{syst}}^2}$$
(6)
is computed, where $`\alpha `$ accounts for possible global normalization effects due to systematic uncertainties (discussed below) across both periods and $`\sigma _{ij}`$ is the statistical uncertainty in each bin. Systematics which can affect spectral shape, mainly energy scale uncertainty, are negligible relative to the statistical uncertainties in the analysis. The function is minimized with respect to $`\alpha `$. The point in the physically allowed parameter space with the smallest chi-square $`\chi _{\mathrm{best}}^2`$ was found, which represents the oscillation scenario best fit by the data.
The 90% confidence level (CL) acceptance region is defined according to the procedure suggested by Feldman and Cousins by:
$$\mathrm{\Delta }\chi ^2=\chi ^2(\mathrm{\Delta }m^2,\mathrm{sin}^22\theta )\chi _{\mathrm{best}}^2>\mathrm{\Delta }\chi _{\mathrm{crit}}^2(\mathrm{\Delta }m^2,\mathrm{sin}^22\theta )$$
(7)
where $`\chi ^2(\mathrm{\Delta }m^2,\mathrm{sin}^22\theta )`$ is the minimized fit quality at the current point in $`\mathrm{\Delta }m^2\mathrm{sin}^22\theta `$ space and $`\mathrm{\Delta }\chi _{\mathrm{crit}}^2`$ is the CL $`\chi ^2`$ cutoff. Due to the sinusoidal dependence of the expected rates on the oscillation parameters and the presence of physically allowed boundaries to those parameters, the cutoff is not simply the $`\mathrm{\Delta }\chi ^2`$ one would analytically find for a three parameter minimization but has to be calculated for each point in the plane. To find the $`\mathrm{\Delta }\chi _{\mathrm{crit}}^2`$ for a point, the experiment is simulated $`10^4`$ times under the assumption that the oscillation hypothesis represented by that point is true. For each simulated data set, a $`\chi _{\mathrm{best}}^2`$ is extracted and a $`\mathrm{\Delta }\chi ^2`$ found for the point. These 10<sup>4</sup> $`\mathrm{\Delta }\chi ^2`$, the simulationsโ fit qualities to the hypothesis, are then ordered. The $`\mathrm{\Delta }\chi ^2`$ of which 90% of the simulations are a better fit is a 90% CL and therefore that oscillation hypothesisโ $`\mathrm{\Delta }\chi _{\mathrm{crit}}^2`$.
The region excluded by the analysis is shown in curve (a) of Fig. 21. The results of the $`\chi ^2`$ analysis, including the oscillation parametersโ best fit to the data, are summarized in the first column of Table VI further below. For the on-off analysis a best fit preferring the no-oscillation hypothesis was found.
In addition to the analysis of the absolute $`\overline{\nu }_\mathrm{e}`$ rates observed, one can analyze the shape of the spectrum of neutrinos seen independently of the absolute normalization, thereby relieving the result of most systematic uncertainties. The $`\chi ^2`$ is calculated at each point in the oscillation parameter plane as in Eqn.( 6), with no constraint on normalization ($`\sigma _{\mathrm{syst}}\mathrm{}`$). The same procedure as before is followed in defining a 90% CL region in the $`\mathrm{sin}^22\theta `$$`\mathrm{\Delta }m^2`$ plane. At large $`\mathrm{\Delta }m^2`$ where $`\overline{\nu }_\mathrm{e}`$ of all energies are oscillating many times within the baseline, the energy spectrum of the incident flux is affected only in magnitude. As a result, the region excluded in the plane does not extend to large $`\mathrm{\Delta }m^2`$, as shown in Fig. 21 (b).
For the spectrum analysis, when the normalization $`\alpha `$ is left free, the minimum $`\chi ^2`$ is obtained for $`\alpha 2`$ (see Table VI) and maximum mixing. This is clearly an unphysical result since such large value of $`\alpha `$ can be excluded to a very high degree of confidence by the independent efficiency calibrations of the detector discussed in previous sections. In addition this result has no effect on the exclusion plot in Fig. 21 because, as shown in Table VI the no-oscillation hypothesis has actually better $`\chi ^2/\mathrm{n}.\mathrm{d}.\mathrm{f}.`$ than the minimum. Also the exclusion plots based on Eqn. (7) and either $`\chi _{\mathrm{best}}^2`$ or $`\chi _{\mathrm{no}.\mathrm{osc}.}^2`$ are found to be virtually identical. Furthermore, changing the bin size from 1 MeV to 0.5 MeV does not appreciably change the exclusion plot, either.
Since the analyses reported above and in the following sections finds no evidence for neutrino oscillations, the spectra of the two years are added and the summed spectrum is plotted in Fig. 22 along with the Monte Carlo expectation.
### B Analysis with the reactor power method
Part of the statistical limitations of the direct subtraction of the preceding analysis is a result of the separation of the data set by year. By correcting the four periods for efficiency and then subtracting the respective reduced flux from full flux periods, the subtraction is forced to treat the $`\overline{\nu }_\mathrm{e}`$ flux from two of the reactors as background. A second $`\chi ^2`$ analysis was performed which effectively uses the full $`\overline{\nu }_\mathrm{e}`$ flux of the refueling periods.
To use the 1998 and 1999 data sets together, the change of both signal and background efficiencies are accounted for. The $`\overline{\nu }_\mathrm{e}`$ efficiency difference is found through the detector Monte Carlo simulation. The efficiency change for background from 1998 to 1999, which is not necessarily the same as for $`\overline{\nu }_\mathrm{e}`$, is extracted via the high statistics fits of correlated background shown in Fig. 19. The 27% increase in background efficiency observed roughly corresponds to what the Monte Carlo simulation predicts for a background composed mainly of double neutron captures. Uncorrelated background accounts for less than 10% of the candidate set; this backgroundโs efficiency changed by a similar amount within the measurement statistics seen in Fig. 16.
The combined data sets are analyzed for oscillation hypotheses by calculating the $`\chi ^2`$ summed over runs $`i`$ (runs with less than ten candidates are combined with adjacent runs):
$$\chi ^2=\underset{i}{}\frac{\left((\alpha N_{\mathrm{calc},i}+b)N_{\mathrm{cand},i}\right)^2}{\sigma _i^2}+\frac{(\alpha 1)^2}{\sigma _{\mathrm{syst}}^2}.$$
(8)
where $`N_{\mathrm{cand},i}`$ is the total $`\overline{\nu }_\mathrm{e}`$ candidate rate, $`N_{\mathrm{calc},i}`$ is the calculated rate, $`\alpha `$ is the overall normalization as before, and $`b`$ is the background rate. The background, $`b`$, is scaled as appropriate for the year but is otherwise assumed to be constant. The function is minimized at each point with respect to $`b`$ and $`\alpha `$. We found no evidence for oscillations and the 90% CL plot, shown in Fig. 23, curve (a), is then constructed around the $`\chi _{\mathrm{best}}^2`$ as before by comparing $`\mathrm{\Delta }\chi ^2`$ with $`\mathrm{\Delta }\chi _{\mathrm{crit}}^2`$ at each point. The predictions of signal and background from this fit for the no-oscillation hypothesis are shown in Table IV. The no-oscillation likelihood and best fit results of the $`\chi ^2`$ analysis with this method are summarized in the third column of Table VI.
### C Analysis with the swap method
A third analysis is used which has the potential of using the full statistical power of the neutrino data set by subtracting background directly. The method, discussed in more detail elsewhere, takes advantage of the asymmetry of the prompt (positron) and delayed (neutron capture) subevents of the neutrino signal. The data selection and trigger treat the two portions of the event identically with the exception of two cuts designed to isolate events with annihilation-like $`\gamma `$โs in the prompt triple.
The candidates remaining after the selection cuts can be written as:
$$N=B_{\mathrm{unc}}+B_{\mathrm{nn}}+B_{\mathrm{pn}}+S_\nu $$
(9)
where $`B_{\mathrm{unc}}`$, $`B_{\mathrm{nn}}`$, and $`B_{\mathrm{pn}}`$ are uncorrelated, two-neutron, and proton-recoilโneutron-capture backgrounds respectively; and $`S_\nu `$ is the neutrino signal. Applying the same neutrino cuts with the positron cuts reversed, or swapped, (such that the positron cuts are now applied to the delayed triple) gives:
$$N^{}=B_{\mathrm{unc}}+B_{\mathrm{nn}}+ฯต_1B_{\mathrm{pn}}+ฯต_2S_\nu $$
(10)
Since the uncorrelated background and two neutron capture backgrounds are symmetric under exchange of the prompt and delayed triples, their efficiencies with the reversed cuts remain the same. The parameters $`ฯต_1`$ and $`ฯต_2`$ denote the relative efficiency change for proton recoils and neutrino signal under the swap, respectively.
The positron cuts are highly efficient for positron annihilation events but have poor efficiency for neutron captures. The Monte Carlo simulation is used to estimate $`ฯต_2=0.159`$. Subtracting (9) from (10) leaves the majority of the neutrino candidates and only proton recoil background:
$$NN^{}=(1ฯต_1)B_{\mathrm{pn}}+(1ฯต_2)S_\nu .$$
(11)
To estimate $`(1ฯต_1)B_{\mathrm{pn}}`$, it is noted that the proton recoil spectrum extends beyond 10 MeV, well above the positron energies of the neutrino signal and other sources of background. These measured high energy events can be used to normalize the $`B_{\mathrm{pn}}`$ background in the signal using the Monte Carlo ratio:
$$r\frac{B_{\mathrm{pn}}^{\mathrm{MC}}(\mathrm{E}_{1,\mathrm{e}^+}<8\mathrm{M}\mathrm{e}\mathrm{V})}{B_{\mathrm{pn}}^{\mathrm{MC}}(\mathrm{E}_{1,\mathrm{e}^+}>10\mathrm{M}\mathrm{e}\mathrm{V})},$$
(12)
where $`B_{\mathrm{pn}}^{\mathrm{MC}}(\mathrm{E}_{1,\mathrm{e}^+}<8\mathrm{M}\mathrm{e}\mathrm{V})`$ is the fraction of simulated $`B_{\mathrm{pn}}`$ events passing the normal $`\overline{\nu }_\mathrm{e}`$ selections, and $`B_{\mathrm{pn}}^{\mathrm{MC}}(\mathrm{E}_{1,\mathrm{e}^+}>10\mathrm{M}\mathrm{e}\mathrm{V})`$ are the fraction of simulated events in the high energy background region. Multiplying the ratio $`r`$ by the measured high energy proton recoil rate gives the $`B_{\mathrm{pn}}`$ background contribution:
$$B_{\mathrm{pn}}=rB_{\mathrm{pn}}^{\mathrm{data}}(\mathrm{E}_{1,\mathrm{e}^+}>10\mathrm{M}\mathrm{e}\mathrm{V}).$$
(13)
The neutrons which cause the proton recoil background are created either by muon capture or spallation in the laboratory walls, or by muons entering the veto counter undetected. The spectrum of the fast neutrons from spallation is not well understood. However, such spectrum can be decoupled somewhat from the resulting proton recoil spectrum. The expected backgrounds were simulated for various possible fast neutron spectra and the resulting $`ฯต_1`$ and $`r`$ for neutrons created in the lab walls were calculated. The same calculation was performed for neutrons created in the passive detector shielding by untagged muons; in this case, the expected yield is much smaller, being only a few percent of that from the walls. The simulated spectra of spallation neutrons are chosen to span the wide range of predictions quoted in literature.
A value for $`ฯต_1`$ of $`1.14\pm 0.07`$ is found after averaging over spectra, implying that the spallation proton recoil background is essentially symmetric like the other backgrounds. Upon simulating the possible spectra, the quantity $`(1ฯต_1)r=0.1\pm 0.05`$ is found to vary little.
The yield and spectrum of neutrons from muon capture are reasonably well understood. Since these neutrons tend to be lower in energy, only those created in the vicinity of the detector have any efficiency for creating background. Knowing the veto inefficiency to miss stopping muons ($`4\pm 1`$)%, the capture rate in water surrounding the detector and its contribution to the background can be estimated using Monte Carlo simulation. Overall this proton recoil background appears to be symmetric as well, $`ฯต_1=0.77\pm 0.32`$, meaning that the subtraction also strongly rejects this background. The uncertainty of the residual background $`(1ฯต_1)B_{\mathrm{pn}}`$ is conservatively estimated to be about 160%, corresponding to $``$4% error on $`N_{\overline{\nu }_\mathrm{e}}`$.
The results of this analysis are summarized in Table V. Overall
$$\frac{R_{\mathrm{obs}}}{R_{\mathrm{calc}}}=1.04\pm 0.03(\mathrm{stat}.)\pm 0.08(\mathrm{sys}.).$$
(14)
The background estimates returned by the reactor power analysis in Table IV compare well with the results of the swap analysis. The 90% CL region for this analysis follows the same $`\chi ^2`$ formula, Eqn. 8, as for the reactor power analysis but uses the background estimated by the swap method subtraction instead of minimizing the function with respect to background. Again, we find no evidence for neutrino oscillations and the excluded region for this analysis is shown in Fig. 23, curve (b).
### D Systematic uncertainties
The systematic uncertainties have three sources: the prediction of expected $`\overline{\nu }_\mathrm{e}`$ interactions, the efficiency estimate, and, for the swap analysis, the $`B_{\mathrm{pn}}`$ estimate. The expected $`\overline{\nu }_\mathrm{e}`$ uncertainty is dominated by the conversion of fission rates into neutrino fluxes, which relies on direct empirical measurements of $`\beta `$ spectra emitted by the isotopes. The Bugey experiment , which directly measured the neutrino flux and energy spectrum emitted by a reactor at short baseline, found agreement within 3% using the same methods; the 3% value is used here as the estimated uncertainty.
The efficiency uncertainty can be further subdivided into that arising from direct comparisons of Monte Carlo e<sup>+</sup> and neutron efficiency from calibration measurements and that arising from the selection cuts themselves. The calibration runs taken with the positron and neutron sources, when compared with Monte Carlo simulations, shows overall agreement across all locations of better than 1% in the efficiency predictions. However, the run-by-run agreement was at a level of 4% for neutrons and 11% for positrons. Since the <sup>22</sup>Na source is similar to the inverse beta decay signal with the e<sup>+</sup> close to detector threshold, the positron efficiency uncertainty over the entire $`\overline{\nu }_\mathrm{e}`$ spectrum was estimated to be closer to 4% in any particular location. These run-by-run variations are then used as our systematic uncertainties in the $`\overline{\nu }_\mathrm{e}`$ efficiency.
To test the robustness of the event selection, each cut is varied within a reasonable range and variations of the ratio between data and Monte Carlo are examined. In order to take into account correlations all cuts were varied simultaneously by randomly sampling a multidimensional cut space. The rms of the resulting ratio of observed/expected is given as the selection cut uncertainty.
The swap method analysis has a somewhat smaller uncertainty for the selection cuts variation as the subtraction tends to cancel out systematics. However, the swap analysis uses a Monte Carlo estimate of the proton recoil background. Due to limited Monte-Carlo statistics and the uncertainty in the fast-neutron energy spectrum, a 4% uncertainty is assigned to the neutrino signal. All of the systematic uncertainties are summarized in Table VII. The total systematic uncertainty is obtained by adding the individual errors in quadrature.
The development of a second simulation and event reconstruction proved to be helpful in understanding systematic uncertainties of the analyses due to the algorithms chosen. For comparison the results for the on-off analysis from both reconstructions are shown in Tables II and III. An independent analysis of systematic errors was performed for the second reconstruction, similar to the method described above, giving comparable results.
## VI Conclusion
In conclusion, the data taken thus far from the Palo Verde experiment show no evidence for $`\overline{\nu }_\mathrm{e}\overline{\nu }_x`$ oscillations. This result, along with the results reported by Chooz and Super-Kamiokande, excludes two family $`\nu _\mu `$$`\nu _\mathrm{e}`$ mixing as being responsible for the atmospheric neutrino anomaly as originally reported by Kamiokande. Later results of Super-Kamiokande, in particular data on the zenith angle distribution of muons and electrons, suggest that muon neutrinos $`\nu _\mu `$ strongly mix with either $`\nu _\tau `$ or with a fourth flavor of neutrino sterile to weak interaction. Clearly it is becoming important to include at least three neutrino flavors when studying results from oscillations experiments.
The most general approach would involve five unknown parameters, three mixing angles and two independent mass differences. However, an intermediate approach consists of a simple generalization of the two flavor scenario, assuming that $`m_3^2m_1^2,m_2^2`$ (i.e. $`\mathrm{\Delta }m_{13}^2=\mathrm{\Delta }m_{23}^2=\mathrm{\Delta }m^2`$, while $`\mathrm{\Delta }m_{12}^20`$). In such a case the mixing angle $`\theta _{12}`$ becomes irrelevant and one is left with only three unknown quantities: $`\mathrm{\Delta }m^2,\theta _{13},\mathrm{and}\theta _{23}`$. With this parameterization the $`\overline{\nu }_e`$ disappearance is governed by
$$P(\overline{\nu }_e\overline{\nu }_x)=\mathrm{sin}^22\theta _{13}\mathrm{sin}^2\frac{\mathrm{\Delta }m^2L}{4E_\nu },$$
(15)
while the $`\nu _\mu \nu _\tau `$ oscillations in this scenario responsible for the atmospheric neutrino results, are described by
$$P(\nu _\mu \nu _\tau )=\mathrm{cos}^4\theta _{13}\mathrm{sin}^22\theta _{23}\mathrm{sin}^2\frac{\mathrm{\Delta }m^2L}{4E_\nu }.$$
(16)
A preliminary analysis of the atmospheric neutrino data based on these assumptions has been performed and its results are shown in Fig. 24 for the $`\overline{\nu }_e`$ disappearance channel. One can see that while the relevant region of the mass difference $`\mathrm{\Delta }m^2`$ is determined by the atmospheric neutrino data, the mixing angle $`\theta _{13}`$ is not constrained very much. Here the reactor neutrino results play a decisive role.
We plan to continue taking data through Summer of 2000, which will provide two additional reduced flux refueling periods.
## Acknowledgments
We would like to thank the Arizona Public Service Company for the generous hospitality provided at the Palo Verde plant. The important contributions of M. Chen, R. Hertenberger, K. Lou, and N. Mascarenhas in the early stages of this project are gratefully acknowledged. We thank K. Scholberg for illuminating discussions on the Super-Kamiokande three flavor analysis. We are indebted to J. Ball, B. Barish, R. Canny, A. Godber, J. Hanson, D. Michael, C. Peck, C. Roat, N. Tolich, and A. Vital for their help. We also acknowledge the generous financial help from the University of Alabama, Arizona State University, California Institute of Technology, and Stanford University. Finally, our gratitude goes to CERN, DESY, FNAL, LANL, LLNL, SLAC, and TJNAF who at different times provided us with parts and equipment needed for the experiment.
This project was supported in part by the Department of Energy. One of us (J.K.) received support from the Hungarian OTKA fund, and another (L.M.) from the ARCS Foundation. |
warning/0003/math0003003.html | ar5iv | text | # Untitled Document
Choix des signes pour la formalitรฉ
de M. Kontsevich
D. Arnal <sup>*</sup><sup>*</sup>Universitรฉ de Metz, Dรฉpartement de Mathรฉmatiques, รฎle du Saulcy, 57045 Metz CEDEX 01. arnal@poncelet.univ-metz.fr, D. Manchon <sup>\**</sup><sup>\**</sup>Institut Elie Cartan, CNRS, BP 239, 54506 Vandลuvre CEDEX. manchon@iecn.u-nancy.fret M. Masmoudi <sup>\***</sup><sup>\***</sup>Universitรฉ de Metz, Dรฉpartement de Mathรฉmatiques, รฎle du Saulcy, 57045 Metz CEDEX 01. masmoudi@poncelet.univ-metz.fr
Rรฉsumรฉ : Lโexpression explicite de la formalitรฉ de M. Kontsevich sur $`\mathrm{}^d`$ est la base de la preuve du thรฉorรจme de formalitรฉ pour une variรฉtรฉ quelconque \[K1 ยง 7\], qui implique ร son tour lโexistence dโรฉtoile-produits sur une variรฉtรฉ de Poisson quelconque. Nous proposons ici un choix cohรฉrent dโorientations et de signes qui permet de reprendre la dรฉmonstration du thรฉorรจme pour $`\mathrm{}^d`$ en tenant compte des signes qui apparaissent devant les diffรฉrents termes de lโรฉquation de formalitรฉ.
Introduction La conjecture de formalitรฉ a รฉtรฉ introduite par M. Kontsevich \[K2\] : elle affirme lโexistence dโun $`L_{\mathrm{}}`$-quasi-isomorphisme de $`\text{g}_1`$ vers $`\text{g}_2`$, oรน $`\text{g}_1`$ et $`\text{g}_2`$ sont les deux algรจbres de Lie diffรฉrentielles graduรฉes naturellement associรฉes ร une variรฉtรฉ $`M`$ : prรฉcisรฉment $`\text{g}_1`$ est lโalgรจbre de Lie diffรฉrentielle graduรฉe des multi-champs de vecteurs munie de la diffรฉrentielle nulle et du crochet de Schouten, et $`\text{g}_2`$ est lโalgรจbre de Lie diffรฉrentielle graduรฉe des opรฉrateurs polydiffรฉrentiels munie de la diffรฉrentielle de Hochschild et du crochet de Gerstenhaber. Les รฉlรฉments de degrรฉ $`n`$ dans $`\text{g}_1`$ sont les $`(n+1)`$-champs de vecteurs, et les รฉlรฉments de degrรฉ $`n`$ dans $`\text{g}_2`$ sont les opรฉrateurs $`(n+1)`$-diffรฉrentiels. Dans les espaces graduรฉs dรฉcalรฉs $`\text{g}_1[1]`$ et $`\text{g}_2[1]`$ ce sont les $`(n+2)`$-champs de vecteurs (resp. les opรฉrateurs $`(n+2)`$-diffรฉrentiels) qui sont de degrรฉ $`n`$. Toute algรจbre de Lie diffรฉrentielle graduรฉe est une $`L_{\mathrm{}}`$-algรจbre. Cela signifie en particulier que les structures dโalgรจbres de Lie diffรฉrentielles graduรฉes sur $`\text{g}_1`$ et $`\text{g}_2`$ induisent des codรฉrivations $`Q`$ et $`Q^{}`$ de degrรฉ $`1`$ sur des cogรจbres $`๐(\text{g}_1)=S^+(\text{g}_1[1])`$ et $`๐(\text{g}_2)=S^+(\text{g}_2[1])`$ respectivement (cf. ยง II.3 et II.4), vรฉrifiant toutes deux lโรฉquation maรฎtresse :
$$[Q,Q]=0,[Q^{},Q^{}]=0.$$
Un $`L_{\mathrm{}}`$-quasi-isomorphisme de $`\text{g}_1`$ vers $`\text{g}_2`$ est par dรฉfinition un morphisme de cogรจbres :
$$๐ฐ:๐(\text{g}_1)๐(\text{g}_2)$$
de degrรฉ zรฉro et commutant aux codรฉrivations, cโest-ร -dire vรฉrifiant lโรฉquation :
$$๐ฐQ=Q^{}๐ฐ,$$
et dont la restriction ร $`\text{g}_1`$ est un quasi-isomorphisme de complexes de $`\text{g}_1`$ dans $`\text{g}_2`$. M. Kontsevich dรฉmontre dans \[K1\] la conjecture de formalitรฉ, cโest-ร -dire lโexistence dโun $`L_{\mathrm{}}`$-quasi-isomorphisme de $`\text{g}_1`$ vers $`\text{g}_2`$, pour toute variรฉtรฉ $`M`$ de classe $`C^{\mathrm{}}`$. La premiรจre รฉtape de la preuve (et mรชme lโessentiel du travail) consiste en la construction explicite du $`L_{\mathrm{}}`$-quasi-isomorphisme $`๐ฐ`$ pour $`M=\mathrm{}^d`$. Le $`L_{\mathrm{}}`$-quasi-isomorphisme $`๐ฐ`$ est uniquement dรฉterminรฉ par ses coefficients de Taylor :
$$๐ฐ_n:S^n(\text{g}_1[1])\text{g}_2[1].$$
(cf. ยง III.2). Si les $`\alpha _k`$ sont des $`s_k`$-champs de vecteurs, ils sont de degrรฉ $`s_k2`$ dans lโespace dรฉcalรฉ $`\text{g}_1[1]`$, et donc $`๐ฐ_n(\alpha _1\mathrm{}\alpha _n)`$ est dโordre $`s_1+\mathrm{}+s_n2n`$ dans $`\text{g}_2[1]`$. Cโest donc un opรฉrateur $`m`$-diffรฉrentiel, avec :
$$\underset{k=1}{\overset{n}{}}s_k=2n+m2.$$
$`()`$
Les coefficients de Taylor sont construits ร lโaide de poids et de graphes : on dรฉsigne par $`G_{n,m}`$ lโensemble des graphes รฉtiquetรฉs et orientรฉs ayant $`n`$ sommets du premier type (sommets aรฉriens) et $`m`$ sommets du deuxiรจme type (sommets terrestres) tels que : 1). Les arรชtes partent toutes des sommets aรฉriens. 2). Le but dโune arรชte est diffรฉrent de sa source (il nโy a pas de boucles). 3). Il nโy a pas dโarรชtes multiples. A tout graphe $`\mathrm{\Gamma }G_{n,m}`$ muni dโun ordre sur lโensemble de ses arรชtes, et ร tout $`n`$-uple de multi-champs de vecteurs $`\alpha _1,\mathrm{},\alpha _n`$ on peut associer de maniรจre naturelle un opรฉrateur $`m`$-diffรฉrentiel $`B_\mathrm{\Gamma }(\alpha _1\mathrm{}\alpha _n)`$ lorsque pour tout $`j\{1,\mathrm{},n\}`$, $`\alpha _j`$ est un $`s_j`$-champ de vecteurs, oรน $`s_j`$ dรฉsigne le nombre dโarรชtes qui partent du sommet aรฉrien numรฉro $`j`$ \[K1 ยง 6.3\]. Le coefficient de Taylor $`๐ฐ_n`$ est alors donnรฉ par la formule :
$$๐ฐ_n(\alpha _1\mathrm{}\alpha _n)=\underset{\mathrm{\Gamma }G_{n,m}}{}W_\mathrm{\Gamma }B_\mathrm{\Gamma }(\alpha _1\mathrm{}\alpha _n),$$
oรน lโentier $`m`$ est reliรฉ ร $`n`$ et aux $`\alpha _j`$ par la formule (\*) ci-dessus. Le poids $`W_\mathrm{\Gamma }`$ est nul sauf si le nombre dโarรชtes $`|E_\mathrm{\Gamma }|`$ du graphe $`\mathrm{\Gamma }`$ est prรฉcisรฉment รฉgal ร $`2n+m2`$. Il sโobtient en intรฉgrant une forme fermรฉe $`\omega _\mathrm{\Gamma }`$ de degrรฉ $`|E_\mathrm{\Gamma }|`$ sur une composante connexe de la compactification de Fulton-McPherson dโun espace de configuration $`C_{A,B}^+`$, qui est prรฉcisรฉment de dimension $`2n+m2`$ \[FM\], \[K1 ยง 5\]. Il dรฉpend lui aussi dโun ordre sur lโensemble des arรชtes, mais le produit $`W_\mathrm{\Gamma }.B_\mathrm{\Gamma }`$ nโen dรฉpend plus. Pour prouver le thรฉorรจme de formalitรฉ M. Kontsevich montre que le morphisme de cogรจbres $`๐ฐ`$ dont les coefficients de Taylor sont les $`๐ฐ_n`$ dรฉfinis ci-dessus est un $`L_{\mathrm{}}`$-quasi-isomorphisme. La mรฉthode consiste ร ramener lโรฉquation de formalitรฉ $`๐ฐQ=Q^{}๐ฐ`$, qui se dรฉveloppe ร lโaide des coefficients de Taylor de $`๐ฐ`$, $`Q`$ et $`Q^{}`$ :
$$\begin{array}{cc}\hfill Q_1^{}๐ฐ_n\left(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n\right)+& \frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\pm Q_2^{}\left(๐ฐ_{|I|}\left(\alpha _I\right)\text{.}๐ฐ_{|J|}\left(\alpha _J\right)\right)=\hfill \\ & =\underset{k=1}{\overset{n}{}}\pm ๐ฐ_n\left(Q_1(\alpha _k)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _k}\text{.}\mathrm{}\text{.}\alpha _n\right)+\hfill \\ & \frac{1}{2}\underset{kl}{}\pm ๐ฐ_{n1}\left(Q_2(\alpha _k\text{.}\alpha _l)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _k}\text{.}\mathrm{}\text{.}\widehat{\alpha _l}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \end{array}$$
ร lโapplication de la formule de Stokes pour les formes $`\omega _\mathrm{\Gamma }`$ sur lโensemble des faces de codimension $`1`$ du bord des espaces de configuration. Nous proposons dans la premiรจre partie de ce travail un choix dโorientation des espaces de configuration (ou plus exactement dโune composante connexe de ceux-ci) $`C_{A,B}^+`$, et un choix cohรฉrent dโorientation pour chacune des faces de codimension $`1`$ du bord de la compactification. Au chapitre II nous explicitons lโisomorphisme $`\mathrm{\Phi }:S^n(\text{g}[1])\stackrel{}{}\mathrm{\Lambda }^n(\text{g})[n]`$ mentionnรฉ dans \[K1 ยง 4.2\] pour tout espace vectoriel graduรฉ g, afin de prรฉciser le passage du langage des algรจbres de Lie diffรฉrentielles graduรฉes et des $`L_{\mathrm{}}`$-algรจbres au langage des $`Q`$-variรฉtรฉs formelles graduรฉes pointรฉes (\[AKSZ\], \[K1 ยง 4.1\]). Nous donnons au chapitre III une formule explicite pour un champ de vecteurs sur une variรฉtรฉ formelle graduรฉe pointรฉe ou un morphisme de variรฉtรฉs formelles graduรฉes pointรฉes en fonction de leurs coefficients de Taylor respectifs. La dรฉmonstration est de nature combinatoire et se fait en explicitant la restriction ร la puissance symรฉtrique $`n`$-iรจme par rรฉcurrence sur $`n`$. Dans le chapitre IV nous exprimons les deux algรจbres de Lie diffรฉrentielles graduรฉes qui nous intรฉressent comme $`Q`$-variรฉtรฉs formelles graduรฉes pointรฉes. Lโisomorphisme dโespaces graduรฉs $`\mathrm{\Phi }`$ explicitรฉ au chapitre II est ici essentiel. Pour la suite nous sommes amenรฉs ร modifier lโalgรจbre de Lie diffรฉrentielle graduรฉe des multi-champs de vecteurs : nous utilisons un crochet de Lie graduรฉ $`[,]^{}`$ liรฉ au crochet de Schouten par la formule :
$$[x,y]^{}=[y,x]_{\text{Schouten}}.$$
Les deux crochets coรฏncident modulo un changement de signe en prรฉsence de deux รฉlรฉments impairs. Nous prรฉcisons au paragraphe IV.4 les signes (du type Quillen) qui apparaissent dans lโรฉquation de formalitรฉ. Enfin nous montrons au chapitre VI que modulo tous les choix effectuรฉs prรฉcรฉdemment $`๐ฐ`$ est bien un $`L_{\mathrm{}}`$-morphisme. Le chapitre V est assez largement indรฉpendant du reste de lโarticle bien que directement reliรฉ ร \[K1\] : nous y donnons une dรฉmonstration dรฉtaillรฉe du thรฉorรจme de quasi-inversion des quasi-isomorphismes donnรฉ dans \[K1 ยง 4.4-4.5\]. Enfin nous rappelons en appendice le lien entre formalitรฉ et quantification par dรฉformation.
I. Orientation des espaces de configuration I.1. Trois choix de paramรฉtrage
Dรฉfinition (Espaces de configuration).
Soit $``$ le demi-plan de Poincarรฉ ($`=\{z\text{},๐ชz>0\}`$). Appelons $`Conf^+(\{z_1,\mathrm{},z_n\};\{t_1,\mathrm{},t_m\})`$ lโensemble des nuages de points :
$$\{(z_1,\mathrm{},z_n;t_1,..,t_m),\text{t. q.}z_i,t_j\text{},z_iz_i^{}\text{si}ii^{},t_1<\mathrm{}<t_m\}$$
et $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ le quotient de cette variรฉtรฉ sous lโaction du groupe $`G`$ de toutes les transformations de la forme :
$$z_iaz_i+b,t_jat_j+b(a>0,b\text{}).$$
La variรฉtรฉ $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ est donc de dimension $`2n+m2`$. Cette variรฉtรฉ est par convention orientรฉe par le passage au quotient de la forme :
$$\mathrm{\Omega }_{\{z_1,\mathrm{},z_n\};\{t_1,\mathrm{},t_m\}}=dx_1dy_1\mathrm{}dx_ndy_ndt_1\mathrm{}dt_m$$
$`z_j=x_j+iy_j`$. Le groupe des transformations considรฉrรฉes prรฉserve lโorientation. On en dรฉduit une orientation des espaces $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$. Plus prรฉcisรฉment, si $`2n+m>0`$, on peut choisir des reprรฉsentants pour paramรฉtrer notre espace. Nous considรฉrons trois mรฉthodes :
Choix 1 :
On choisit lโun des $`z_i`$ (disons $`z_{j_0}=x_{j_0}+iy_{j_0}`$) et on le place au point $`i`$ par une transformation de $`G`$. Les autres points sont alors fixรฉs :
$$p_{j_0}=i,p_j=\frac{z_jx_{j_0}}{y_{j_0}},q_l=\frac{t_lx_{j_0}}{y_{j_0}}.$$
Dans ce cas, on paramรจtre $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par les coordonnรฉes des $`p_j=a_j+ib_j`$ ($`jj_0`$) et les $`q_l`$, lโorientation, dans ces coordonnรฉes de $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ est celle donnรฉe par la forme :
$$\mathrm{\Omega }=\underset{jj_0}{}(da_jdb_j)dq_1\mathrm{}dq_m.$$
(Lโordre sur les indices $`j`$ nโimporte pas car les 2-formes $`da_jdb_j`$ commutent entre elles).
Choix 2 :
On choisit lโun des $`t_l`$ (disons $`t_{l_0}`$) et on le place en 0 par une translation, puis on fait une dilatation pour forcer le module de lโun des $`z_j`$ (disons $`z_{j_0}`$) ร valoir 1 :
$$p_{j_0}=\frac{z_{j_0}t_{l_0}}{|z_{j_0}t_{j_0}|}=e^{i\theta _{j_0}},p_j=\frac{z_jt_{l_0}}{|z_{j_0}t_{j_0}|},q_{l_0}=0,q_l=\frac{t_lt_{l_0}}{|z_{j_0}t_{j_0}|}.$$
On paramรจtre alors $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par lโargument $`\theta _{j_0}`$ de $`p_{j_0}`$ (compris entre 0 et $`\pi `$) et par les coordonnรฉes des $`p_j`$ ($`jj_0`$) et les $`q_l`$ ($`ll_0`$). Lโorientation, dans ces coordonnรฉes de $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ est celle donnรฉe par la forme :
$$\mathrm{\Omega }=(1)^{l_01}d\theta _{j_0}\underset{jj_0}{}(da_jdb_j)dq_1\mathrm{}\widehat{dq_{l_0}}\mathrm{}dq_m.$$
En effet, on part de la forme $`\mathrm{\Omega }`$ du cas 1, avec $`j_0=1`$, puisque lโordre des $`p`$ nโintervient pas, on place $`q_{l_0}`$ โen tรชteโ :
$$\mathrm{\Omega }=(1)^{l_01}dq_{l_0}da_2db_2\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_{l_0}}\mathrm{}dq_{l_1}\mathrm{}dq_n,$$
puis on effectue le changement de variables :
$$p_1^{}=\frac{iq_{l_0}}{|iq_{l_0}|}=e^{i\theta _1},p_j^{}=\frac{p_jq_{l_0}}{|iq_{l_0}|}=a_j^{}+ib_j^{}(2jn),q_k^{}=\frac{q_kq_{l_0}}{|iq_{l_0}|}(kl_0).$$
Dont le jacobien $`\frac{1}{(1+q_{l_0}^2)^{\frac{m+n}{2}}}`$ est strictement positif, pour obtenir la forme annoncรฉe.
Choix 3 :
On choisit deux points $`t_{l_0}<t_{l_1}`$, on amรจne par une translation le premier en $`0`$ et le second en 1 par une dilatation.
$$p_j=\frac{z_jt_{l_0}}{t_{l_1}t_{l_0}},q_{l_0}=0,q_{l_1}=1,q_l=\frac{t_lt_{l_0}}{t_{l_1}t_{l_0}}.$$
On paramรจtre $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par les coordonnรฉes des $`p_j=a_j+ib_j`$ et par les $`q_l`$ ($`ll_0`$ et $`ll_1`$). lโorientation est donnรฉe par la forme :
$$\mathrm{\Omega }=(1)^{l_0+l_1+1}\underset{j=1}{\overset{n}{}}(da_jdb_j)dq_1\mathrm{}\widehat{dq_{l_0}}\mathrm{}\widehat{dq_{l_1}}\mathrm{}dq_m.$$
En effet, on part de la forme $`\mathrm{\Omega }`$ du cas 1, on place $`q_{l_0}`$ et $`q_{l_1}`$ โen tรชteโ :
$$\mathrm{\Omega }=(1)^{l_01+l_12}dq_{l_0}dq_{l_1}da_2db_2\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_{l_0}}\mathrm{}dq_n,$$
puis on effectue le changement de variables :
$$p_1^{}=\frac{iq_{l_0}}{q_{l_1}q_{l_0}}=a_1+ib_1,p_j^{}=\frac{p_jq_{l_0}}{q_{l_1}q_{l_0}}=a_j^{}+ib_j^{}(2jn),q^{}k=\frac{q_kq_{l_0}}{q_{l_1}q_{l_0}}(kl_0,kl_1).$$
Dont le jacobien $`\frac{q_{l_1}q_{l_0}}{(q_{l_1}q_{l_0})^{1+n+m}}`$ est strictement positif, pour obtenir la forme annoncรฉe.
I.2. Compactification des espaces de configuration
On plonge lโespace de configuration $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ dans une variรฉtรฉ compacte de la faรงon suivante. Chaque fois que lโon prend deux points $`A`$ et $`B`$ du nuage de points $`(z_j,\overline{z_j};t_l)`$, on leur associe lโangle $`Arg(BA)`$, ร chaque triplet de points $`(A,B,C)`$ du nuage, on associe lโรฉlรฉment $`[AB,BC,CA]`$ de lโespace projectif $`\text{}^2(\text{})`$ quโils dรฉfinissent. On a ainsi une application :
$$\stackrel{~}{\mathrm{\Phi }}:Conf^+(z_j,t_l)\text{๐}^{(2n+m)(2n+m1)}\times \left(\text{}^2(\text{})\right)^{(2n+m)(2n+m1)(2n+m2)}.$$
Cette application passe au quotient et il nโest pas difficile de montrer que lโon obtient ainsi un plongement
$$\mathrm{\Phi }:C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+\text{๐}^{(2n+m)(2n+m1)}\times \left(\text{}^2(\text{})\right)^{(2n+m)(2n+m1)(2n+m2)}.$$
On dรฉfinit la compactification $`\overline{C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+}`$ de $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ comme รฉtant la fermeture dans $`\text{๐}^{(2n+m)(2n+m1)}\times \left(\text{}^2(\text{})\right)^{(2n+m)(2n+m1)(2n+m2)}`$ de $`\mathrm{\Phi }\left(C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+\right)`$. On obtient ainsi une variรฉtรฉ ร coins et on cherche son bord $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$.
Les points du bord sโobtiennent par une succession de collapses de points du nuage. On retrouve la description de M. Kontsevich ร deux dรฉtails prรจs : lorsque des points aรฉriens (cโest ร dire un ou des $`p_j`$) se rapprochent de , il faut distinguer entre quels $`q_l`$ ils arrivent, il y a trop de faces du bord, puisque les faces correspondant au rapprochement de points terrestres (des $`q_l`$) non contigus est impossible sans que tous les points qui les sรฉparent se rapprochent aussi. En codimension 1, on obtient deux types de faces : I.2.1. Faces de type 1 Parmi les points aรฉriens, $`n_1`$ points se rapprochent en un point $`p`$ qui reste aรฉrien. Une telle face existe si $`nn_12`$. A la limite, on obtient une variรฉtรฉ produit :
$$F=_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\}}C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+=C_{\{p_{i_1},\mathrm{},i_{n_1}\}}\times C_{\{p,p_1,\mathrm{},\widehat{p_{i_1}},\mathrm{},\widehat{p_{i_{n_1}}},\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}$$
$`()`$
oรน lโespace $`C_{\{p_1,\mathrm{},p_{n_1}\}}`$ est le quotient de lโespace $`Conf(z_1,\mathrm{},z_{n_1})`$ par lโaction du groupe $`G^{}`$ des transformations $`z_jaz_j+b`$ ($`a>0`$ et $`b\text{}`$). cโest une varitฬรฉ de dimension $`2n_13`$ ($`n_12`$). On la plonge dans un produit de tores et dโespaces projectifs comme pour $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$. Enfin on lโoriente de la faรงon suivante; $`z_1`$ est placรฉ en 0 par une translation complexe puis $`|z_2|`$ est normalisรฉ ร 1 par une dilatation,
$$p_1=0,p_2=\frac{z_2z_1}{|z_2z_1|}=e^{i\theta _2},p_j=\frac{z_jz_1}{|z_2z_1|}=a_j+ib_j$$
et on prend lโorientation dรฉfinie par la forme :
$$\mathrm{\Omega }_1=d\theta _2\underset{j3}{}(da_jdb_j).$$
Orientons maintenant la face $`F`$. On choisit la forme volume $`\mathrm{\Omega }_1\mathrm{\Omega }_2`$ sur le produit $`()`$$`\mathrm{\Omega }_2`$ est lโune des formes dรฉfinies ci-dessus pour orienter $`C_{\{p_j\};\{q_l\}}^+`$. Lโorientation de la face ร partir de celle de $`\mathrm{\Omega }`$ est $`\pm \mathrm{\Omega }_1\mathrm{\Omega }_2`$.
Lemme I.2.1.
La face $`F`$ est orientรฉe par $`\mathrm{\Omega }_F=\mathrm{\Omega }_1\mathrm{\Omega }_2`$.
Dรฉmonstration. On a vu que lโon pouvait changer lโordre des points $`p_j`$ de $`C_{\{p_j\};\{q_l\}}^+`$ sans changer lโorientation. On renumรฉrote les points $`p_{i_1}`$,โฆ, $`p_{i_{n_1}}`$ en $`p_1`$, $`p_2`$,โฆ, $`p_{n_1}`$, puis on fixe $`p_1=i`$ :
$$\mathrm{\Omega }=\underset{j=2}{\overset{n_1}{}}(da_jdb_j)\mathrm{\Omega }_2,$$
ensuite on change de variables dans le premier facteur en posant :
$$p_2^{}=e^{i\theta _2},p_j^{}=\frac{p_j}{|p_2i|}=a_j^{}+ib_j^{}(j=3,\mathrm{},n_1).$$
Lorsque les $`n_1`$ premiers points collapsent, on agrandit le petit nuage quโils forment en normalisant la distance qui sรฉpare les 2 premiers ร 1. Posons $`\rho _2=|p_2i|`$. le changement de variable donne pour $`\mathrm{\Omega }`$ la forme :
$$\mathrm{\Omega }^{}=d\rho _2d\theta _2\underset{j3}{}(da_j^{}db_j^{})\mathrm{\Omega }_2.$$
La face est obtenue lorsque $`\rho _20`$. Or $`\rho _2>0`$, on doit donc lโorienter avec
$$\mathrm{\Omega }_F=d\theta _2\underset{j3}{}(da_j^{}db_j^{})\mathrm{\Omega }_2=\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
$``$
I.2.2. Face de type 2 Parmi les points du nuage, $`n_1`$ points aรฉriens et $`m_1`$ points terrestres se rapprochent en un point $`q`$ terrestre. Une telle face existe si $`n+m>n_1+m_1`$ et $`2n_1+m_12`$. A la limite, on obtient une variรฉtรฉ produit :
$$\begin{array}{cc}\hfill F& =_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_{l+1},\mathrm{},q_{l+m_1}\}}C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+\hfill \\ & =C_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_{l+1},\mathrm{},q_{l+m_1}\}}\times C_{\{p_1,\mathrm{},\widehat{p_{i_1}},\mathrm{},\widehat{p_{i_{n_1}}},\mathrm{},p_n\};\{q_1,\mathrm{},q_l,q,q_{l+m_1+1},\mathrm{},q_m\}}.\hfill \end{array}$$
$`()`$
On appelle $`\mathrm{\Omega }_1`$ et $`\mathrm{\Omega }_2`$ lโune des formes volumes de chacun des facteurs de ce produit. La forme $`\mathrm{\Omega }_1\mathrm{\Omega }_2`$ est une forme volume sur $`F`$. On donne lโorientation de $`F`$ ร partir de celle de lโespace de configuration de dรฉpart en terme de cette forme.
Lemme I.2.2.
Avec nos notations, la face $`F`$ est orientรฉe par :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Dรฉmonstration. Il faut considรฉrer six types de nuages diffรฉrents :
Sous-cas 1 : $`n>n_1>0`$
On suppose que $`p_1`$,โฆ, $`p_{n_1}`$ et $`q_{l+1}`$, โฆ, $`q_{l+m_1}`$ collapsent. On paramรจtre lโespace $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par $`p_{n_1+1}=i`$. La forme dโorientation est :
$$\begin{array}{cc}\hfill \mathrm{\Omega }=& (1)^{lm_1}da_1db_1\mathrm{}da_{n_1}db_{n_1}dq_{l+1}\mathrm{}dq_{l+m_1}\hfill \\ & da_{n_1+2}db_{n_1+2}\mathrm{}da_ndb_ndq_1\mathrm{}dq_ldq_{l+m_1+1}\mathrm{}dq_m.\hfill \\ \hfill =& (1)^{lm_1+l+m_1+1}db_1\mathrm{}da_{n_1}db_{n_1}dq_{l+1}\mathrm{}dq_{l+m_1}\hfill \\ & da_{n_1+2}db_{n_1+2}\mathrm{}da_ndb_ndq_1\mathrm{}dq_lda_1dq_{l+m_1+1}\mathrm{}dq_m.\hfill \end{array}$$
(Certains termes peuvent ne pas apparaรฎtre, par exemple si $`n=n_1+1`$ ou $`m=m_1`$). On change de variables en posant :
$$a_1=q,p_j^{}=\frac{p_ja_1}{b_1}(2jn_1),q_k^{}=\frac{q_ka_1}{b_1}(l+1kl+m_1).$$
Alors on peut รฉcrire de faรงon un peu abusive :
$$\mathrm{\Omega }=(1)^{lm_1+l+m_1+1}db_1\mathrm{\Omega }_1\mathrm{\Omega }_2$$
et puisque $`b_1>0`$ et la face $`F`$ est obtenue pour $`b_1=0`$, son orientation est donnรฉe par :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Sous-cas 2 : $`n=n_1>0`$ (et donc $`mm_1+1`$) et $`l>0`$
On suppose que $`p_1`$,โฆ, $`p_n`$ et $`q_{l+1}`$, โฆ, $`q_{l+m_1}`$ collapsent. On paramรจtre lโespace $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par $`q_l=0`$, $`q_{l+1}=1`$. La forme dโorientation est :
$$\begin{array}{cc}\hfill \mathrm{\Omega }& =(1)^{2l+2}da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_l}\widehat{dq_{l+1}}\mathrm{}dq_m\hfill \\ & =(1)^{(l1)(m_11)}da_1db_1\mathrm{}da_ndb_ndq_{l+2}\mathrm{}dq_{l+m_1}\hfill \\ & dq_1\mathrm{}dq_{l1}dq_{l+m_1+1}\mathrm{}dq_m\hfill \end{array}.$$
On change de variables en posant :
$$a_1+ib_11=\rho _1e^{i\theta _1}p_j^{}=\frac{p_j1}{\rho _1}(2jn),q_k^{}=\frac{q_k1}{\rho _1}(l+2kl+m_1).$$
Alors
$$\begin{array}{cc}\hfill \mathrm{\Omega }& =(1)^{lm_1+l+m_1+1}\rho _1^{(n+m_15)}d\rho _1d\theta _1da_2db_2\mathrm{}da_ndb_ndq_{l+2}\mathrm{}dq_{l+m_1}\hfill \\ & dq_1\mathrm{}dq_{l1}dq_{l+m_1+1}\mathrm{}dq_m\hfill \end{array}.$$
On peut donc รฉcrire de faรงon un peu abusive :
$$\mathrm{\Omega }(1)^{lm_1+l+m_1+1}d\rho _1\mathrm{\Omega }_1(1)^{l1+l+12}\mathrm{\Omega }_2$$
et puisque $`\rho _1>0`$ et la face $`F`$ est obtenue pour $`\rho _1=0`$, son orientation est donnรฉe par :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Sous-cas 3 : $`n=n_1>0`$ (et donc $`mm_1+1`$) et $`l=0`$
Cโest le mรชme calcul que ci-dessus, on pose $`q_{m_1}=0`$, $`q_{m_1+1}=1`$, on obtient :
$$\mathrm{\Omega }=(1)^{2m_1+2}da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_{m_1}}\widehat{dq_{m_1+1}}\mathrm{}dq_m.$$
On change de variables en posant :
$$a_1+ib_1=\rho _1e^{i\theta _1}p_j^{}=\frac{p_j}{\rho _1}(2jn),q_k^{}=\frac{q_k}{\rho _1}(1km_11).$$
Alors
$$\begin{array}{cc}\hfill \mathrm{\Omega }& =\rho _1d\rho _1d\theta _1da_2db_2\mathrm{}da_ndb_ndq_1\mathrm{}dq_{m_11}dq_{m_1+2}\mathrm{}dq_m\hfill \\ & d\rho _1(1)^{m_11}\mathrm{\Omega }_1(1)^{11+22}\mathrm{\Omega }_2\hfill \end{array}$$
et puisque $`\rho _1>0`$ et la face $`F`$ est obtenue pour $`\rho _1=0`$, son orientation est donnรฉe par :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Sous-cas 4 : $`nn_1=0`$ (et donc $`m_1>1`$) et $`l>0`$
On suppose que $`q_{l+1}`$, โฆ, $`q_{l+m_1}`$ collapsent. On paramรจtre lโespace $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ par $`q_l=0`$, $`q_{l+1}=1`$. La forme dโorientation est :
$$\begin{array}{cc}\hfill \mathrm{\Omega }& =(1)^{2l+2}da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_l}\widehat{dq_{l+1}}\mathrm{}dq_m\hfill \\ & =(1)^{(l1)(m_11)}dq_{l+2}\mathrm{}dq_{l+m_1}\hfill \\ & da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}dq_{l1}dq_{l+m_1+1}\mathrm{}dq_m\hfill \end{array}.$$
On change de variables en posant :
$$q_k^{}=\frac{q_kq_{l+2}}{q_{l+2}1}(l+3kl+m_1).$$
Alors :
$$\begin{array}{cc}\hfill \mathrm{\Omega }& (1)^{lm_1+l+m_1+1}dq_{l+2}dq_{l+3}^{}\mathrm{}dq_{l+m_1}^{}\hfill \\ & da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}dq_{l1}dq_{l+m_1+1}\mathrm{}dq_m\hfill \end{array}.$$
On peut donc รฉcrire de faรงon un peu abusive :
$$\mathrm{\Omega }=(1)^{lm_1+l+m_1+1}dq_{l+2}\mathrm{\Omega }_1(1)^{l1+l+12}\mathrm{\Omega }_2$$
et puisque $`q_{l+2}1>0`$ et la face $`F`$ est obtenue pour $`q_{l+2}1=0`$, son orientation est donnรฉe par :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Sous-cas 5 : $`nn_1=0`$ (et donc $`m_1>1`$), $`l=0`$ et $`m_1<m`$
On pose $`q_{m_1}=0`$, $`q_{m_1+1}=1`$, on obtient :
$$\mathrm{\Omega }=(1)^{2m_1+2}da_1db_1\mathrm{}da_ndb_ndq_1\mathrm{}\widehat{dq_{m_1}}\widehat{dq_{m_1+1}}\mathrm{}dq_m.$$
On change de variables en posant :
$$q_k^{}=\frac{q_kq_{m_11}}{q_{m_11}}(1km_12)$$
Alors :
$$\begin{array}{cc}\hfill \mathrm{\Omega }& dq_1^{}\mathrm{}dq_{m_12}^{}dq_{m_11}da_1db_1\mathrm{}da_ndb_ndq_{m_1+2}\mathrm{}dq_m\hfill \\ & =(1)^{m_1}dq_{m_11}\mathrm{\Omega }_1(1)^{1+21}\mathrm{\Omega }_2.\hfill \end{array}$$
Maintenant $`q_{m_11}<0`$ et lโorientation de la face est encore :
$$\mathrm{\Omega }_F=(1)^{lm_1+m_1+l}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Sous-cas 6 : $`nn_1=0`$ (et donc $`m_1>1`$), $`l=0`$ et $`m_1=m`$ et donc $`n>0`$
On pose $`q_1=0`$ et $`p_1=e^{i\theta _1}`$. La forme $`\mathrm{\Omega }`$ est
$$\begin{array}{cc}\hfill \mathrm{\Omega }& =d\theta _1da_2db_2\mathrm{}da_ndb_ndq_2\mathrm{}dq_m\hfill \\ & =(1)^{m1}dq_2\mathrm{}dq_md\theta _1da_2db_2\mathrm{}da_ndb_n.\hfill \end{array}$$
On change de variables en posant :
$$q_k^{}=\frac{q_kq_2}{q_2}(3km)$$
Alors :
$$\mathrm{\Omega }(1)^{m1}dq_2(1)^{11+22}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Puisque $`q_2>0`$ et $`F`$ apparaรฎt pour $`q_2=0`$, lโorientation de la face est encore :
$$\mathrm{\Omega }_F=(1)^{lm_1+l+m_1}\mathrm{\Omega }_1\mathrm{\Omega }_2.$$
Tout nuage de point correspondant ร une face de type 2 relรจve dโun de ces six sous-cas. Ceci termine la dรฉmonstration du lemme I.2.2.
$``$
II. Algรจbres symรฉtriques et extรฉrieures sur les espaces graduรฉs II.1. La catรฉgorie des espaces graduรฉs Un espace vectoriel sur un corps $`k`$ est graduรฉ sโil est muni dโune $`\mathrm{}`$graduation :
$$V=\underset{n\mathrm{}}{}V_n$$
Le degrรฉ dโun รฉlรฉment homogรจne $`x`$ sera notรฉ $`|x|`$. Un espace graduรฉ sera toujours considรฉrรฉ comme un super-espace vectoriel, la $`\mathrm{}_2`$graduation รฉtant dรฉduite de la $`\mathrm{}`$graduation :
$$V_+=\underset{n\mathrm{}}{}V_{2n}V_{}=\underset{n\mathrm{}}{}V_{2n+1}$$
Si $`V`$ et $`W`$ sont des espaces graduรฉs, il existe une graduation naturelle sur $`VW`$, $`VW`$, $`\text{Hom}_k(V,W)`$. Un morphisme dโespaces graduรฉs entre $`V`$ et $`W`$ est par dรฉfinition un รฉlรฉment de degrรฉ zรฉro dans $`\text{Hom}_k(V,W)`$. Une algรจbre graduรฉe est un espace graduรฉ $`B`$ muni dโune structure dโalgรจbre telle que la multiplication $`m:BBB`$ est un morphisme dโespaces graduรฉs, cโest-ร -dire :
$$B_iB_jB_{i+j}$$
On dรฉfinit de la mรชme maniรจre les notions de $`B`$ modules graduรฉs ร gauche ou ร droite. Si $`A`$ et $`B`$ sont deux algรจbres graduรฉes, le produit :
$$\begin{array}{cc}\hfill m_{AB}:ABAB& AB\hfill \\ \hfill aba^{}b^{}& (1)^{|b||a^{}|}aa^{}bb^{}\hfill \end{array}$$
est associatif et munit $`AB`$ dโune structure dโalgรจbre graduรฉe. Si $`M`$ (resp. $`N`$) est un $`A`$-module (resp. un $`B`$-module) graduรฉ ร gauche, la mรชme rรจgle des signes (la rรจgle de Koszul) permet de dรฉfinir une structure de $`AB`$module graduรฉ ร gauche sur $`MN`$. Si $`A,A^{},B,B^{}`$ sont des espaces graduรฉs, lโidentification de $`\text{Hom}_k(AB,A^{}B^{})`$ avec $`\text{Hom}_k(A,A^{})\text{Hom}_k(B,B^{})`$ se fait avec la mรชme rรจgle des signes :
$$(fg)(ab)=(1)^{|g||a|}f(a)g(b)$$
Une algรจbre graduรฉe est dite commutative si on a :
$$xy(1)^{|x||y|}yx=0$$
Une cogรจbre graduรฉe $`C`$ se dรฉfinit de maniรจre similaire : la comutiplication doit vรฉrifier :
$$\mathrm{\Delta }C_j\underset{k+l=j}{}C_kC_l$$
On dรฉfinit une structure de cogรจbre graduรฉe sur le produit tensoriel de deux cogรจbres graduรฉes en appliquant la mรชme rรจgle sur les signes que dans le cas des algรจbres. Une dรฉrivation de degrรฉ $`i`$ dans une algรจbre graduรฉe $`B`$ est un morphisme linรฉaire $`d:BB`$ de degrรฉ $`i`$ tel que :
$$d(xy)=dx.y+(1)^{i|x|}x.dy$$
ce qui sโรฉcrit encore :
$$dm=m(dI+Id)$$
$`m`$ dรฉsigne la multiplication de lโalgรจbre (attention ร la rรจgle des signes). Une codรฉrivation de degrรฉ $`i`$ dans une cogรจbre graduรฉe $`C`$ est un morphisme linรฉaire $`d:CC`$ de degrรฉ $`i`$ tel que si $`\mathrm{\Delta }X=_{(x)}x^{}x^{\prime \prime }`$ on a :
$$\mathrm{\Delta }dx=\underset{(x)}{}dx^{}x^{\prime \prime }+(1)^{i|x^{}|}x^{}dx^{\prime \prime }$$
ou encore :
$$\mathrm{\Delta }d=(dI+Id)\mathrm{\Delta }$$
Enfin une algรจbre de Lie graduรฉe est un espace vectoriel graduรฉ g muni dโun crochet $`[.,.]`$ tel que : 1) $`[\text{g}_i,\text{g}_j]\text{g}_{i+j}`$ 2) $`[x,y]=(1)^{|x||y|}[y,x]`$ 3) $`(1)^{|x||z|}[[x,y],z]+(1)^{|y||x|}[[y,z],x]+(1)^{|z||y|}[[z,x],y]=0`$
(identitรฉ de Jacobi graduรฉe). Lโidentitรฉ de Jacobi graduรฉe sโexprime aussi en disant que $`\text{ad}x=[x,.]`$ est une dรฉrivation (de degrรฉ $`|x|`$). Une algรจbre de lie graduรฉe est diffรฉrentielle si elle est munie dโune diffรฉrentielle $`d`$ de degrรฉ 1 ($`d:๐ค๐ค[1]`$), telle que :
$$d^2=0,d\left([x,y]\right)=[dx,y]+(1)^{1.|x|}[x,dy].$$
II.2 La rรจgle de Koszul La raison profonde qui fait que โla rรจgle des signes marcheโ est la suivante : la catรฉgorie des espaces vectoriels $`\mathrm{}_2`$graduรฉs munie du produit tensoriel $``$ usuel et des applications :
$$\begin{array}{cc}\hfill \tau _{A,B}:AB& BA\hfill \\ \hfill ab& (1)^{|a||b|}ba\hfill \end{array}$$
est une catรฉgorie tensorielle tressรฉe, cโest ร dire que les tressages $`\tau _{A,B}`$ sont fonctoriels :
$$\begin{array}{ccc}AB& \underset{\tau _{A,B}}{}& BA\\ {}_{}{}^{fg}& & ^{gf}\\ A^{}B^{}& \underset{\tau _{A^{},B^{}}}{}& B^{}A^{}\end{array}$$
et vรฉrifient :
$$\tau _{AB,C}=(\tau _{A,C}I_B)(I_A\tau _{B,C})$$
De ces deux propriรฉtรฉs on dรฉduit facilement lโรฉquation de lโhexagone, cโest-ร -dire la commutativitรฉ du diagramme suivant :
$$\begin{array}{ccc}& ABC& \\ & & \\ ACB& & BAC\\ & & \\ CAB& & BCA\\ & & \\ & CBA& \end{array}$$
La catรฉgorie tensorielle tressรฉe des espaces $`\mathrm{}_2`$graduรฉs peut aussi se voir comme la catรฉgorie des modules sur lโalgรจbre de Hopf quasi-triangulaire $`(H_2,R)`$$`H_2`$ est lโalgรจbre du groupe $`\mathrm{}_2`$ munie de la multiplication et de la comultiplication usuelle, mais oรน la $`R`$matrice est non triviale. Dans cette catรฉgorie le carrรฉ des tressages est toujours lโidentitรฉ (cโest une catรฉgorie tensorielle stricte). On peut faire de mรชme avec des espaces $`\mathrm{}_k`$graduรฉs en remplaรงant $`1`$ par $`e^{\frac{2i\pi }{k}}`$. On obtient ainsi la catรฉgorie des espaces vectoriels anyoniques, qui est tressรฉe de maniรจre effective pour $`k3`$ \[M\]. II.3. Dรฉcalages Soit $`V`$ un espace graduรฉ. On pose :
$$V[1]=Vk[1]$$
$`k[1]`$ est lโespace graduรฉ tel que $`k_n=\{0\}`$ pour $`n1`$ et $`k_1=k`$. Autrement dit $`V[1]`$ et $`V`$ ont mรชme espace vectoriel sous-jacent, mais le degrรฉ dโun รฉlรฉment est baissรฉ dโune unitรฉ dans $`V[1]`$. On posera en outre :
$$[n]=[1]^n$$
pour tout entier $`n`$. II.4. Algรจbres symรฉtriques et extรฉrieures Lโalgรจbre symรฉtrique $`S(V)`$ (resp. lโalgรจbre extรฉrieure $`\mathrm{\Lambda }(V)`$) est dรฉfinie par :
$$S(V)=T(V)/<xy(1)^{|x||y|}yx>\text{ resp.}\mathrm{\Lambda }(V)=T(V)/<xy+(1)^{|x||y|}yx>$$
Ce sont des espaces graduรฉs de maniรจre naturelle. La proposition suivante est implicite dans \[K1\] :
Proposition II.4.1 (symรฉtrisation).
Pour tout espace vectoriel graduรฉ $`V`$ et pour tout $`n>0`$ on a un isomorphisme naturel :
$$\mathrm{\Phi }_n:S^n(V[1])\stackrel{~}{}\mathrm{\Lambda }^n(V)[n]$$
donnรฉ par :
$$\mathrm{\Phi }_n(x_1.\mathrm{}.x_n)=\alpha (x_1,\mathrm{},x_n)x_1\mathrm{}x_n$$
oรน, pour des $`x_i`$ homogรจnes, $`\alpha (x_1,\mathrm{},x_n)`$ dรฉsigne la signature de la permutation โunshuffleโ qui range les $`x_i`$ pairs dans $`V`$ ร gauche sans les permuter, et les $`x_i`$ impairs dans $`V`$ ร droite sans les permuter.
Dรฉmonstration. Soit $`I`$ (resp. $`J`$) lโensemble des $`i`$ tels que $`x_i`$ soit de degrรฉ pair (resp. impair), et $`\alpha (I,J)=\alpha (x_1,\mathrm{},x_n)`$ la signature de la permutation-rangement associรฉe. Lโisomorphisme $`\mathrm{\Phi }_n`$ est donnรฉ par la restriction ร $`S^n(V[1])`$ de la composition des trois flรจches du diagramme ci-dessous (la flรจche supรฉrieure est un isomorphisme dโalgรจbres) :
$$\begin{array}{ccc}S(V[1])& \stackrel{~}{}& S(V[1]_+)S(V[1]_{})\\ x_1\mathrm{}x_n& & x_Ix_J\\ & & \\ \mathrm{\Lambda }(V)& \stackrel{~}{}& \mathrm{\Lambda }(V_{})\mathrm{\Lambda }(V_+)\\ \alpha (I,J)x_1\mathrm{}x_n& & x_Ix_J\end{array}$$
Enfin si les $`x_j`$ sont de degrรฉ $`d_j`$ dans $`V[1]`$, $`x_1\mathrm{}x_n`$ est de degrรฉ $`d_1+\mathrm{}+d_n`$ dans $`S^n(V[1])`$, donc de degrรฉ $`d_1+\mathrm{}+d_n+n`$ dans $`S^n(V)`$. $`\mathrm{\Phi }_n(x_1\mathrm{}x_n)`$ est donc de degrรฉ $`d_1+\mathrm{}+d_n+n`$ dans $`\mathrm{\Lambda }^n(V)`$, donc de degrรฉ $`d_1+\mathrm{}+d_n`$ dans $`\mathrm{\Lambda }^n(V)[n].`$
$``$
Remarque : Lโapplication $`\mathrm{\Phi }=\mathrm{\Phi }_n`$ est un morphisme dโespace vectoriel graduรฉ mais pas dโalgรจbre. Il est dโailleurs vain de vouloir chercher un isomorphisme dโalgรจbres entre $`S(V[1])`$ et $`\mathrm{\Lambda }^n(V)[n]`$, car deux รฉlรฉments de paritรฉ opposรฉe commutent dans le premier cas, et anticommutent dans le second cas. II.5. Un exemple : $`Tens(\mathrm{}^d)`$ Lโalgรจbre des tenseurs contravariants totalement antisymรฉtriques est une algรจbre naturellement graduรฉe par lโordre des tenseurs. On aimerait la voir comme lโespace sous-jacent ร une algรจbre symรฉtrique. Notons donc $`V`$ lโespace vectoriel $`๐ณ\left(\text{}^d\right)`$ des champs de vecteurs sur $`\text{}^d`$, graduรฉ par $`V=V_0`$. On identifie $`Tens_n\left(\text{}^d\right)=^nV`$ ร $`S^n\left(V[1]\right)[n]`$ par $`\mathrm{\Phi }_n`$.
Dans la suite, on posera
$$T_{poly}\left(\text{}^d\right)=Tens\left(\text{}^d\right)[1].$$
III. Variรฉtรฉs formelles graduรฉes III.1. Variรฉtรฉs formelles On se place sur le corps des rรฉels ou des complexes. On se donne un voisinage ouvert $`U`$ de $`0`$ dans $`\mathrm{}^d`$. Une fonction analytique $`\phi `$ sur $`U`$ ร valeurs dans $`\mathrm{}`$ est dรฉterminรฉe par son dรฉveloppement de Taylor en $`0`$ :
$$\phi (x)=\underset{\alpha \mathrm{}^d}{}\frac{x^\alpha }{\alpha !}(^\alpha \phi )(0)$$
On รฉtablit ainsi une dualitรฉ non dรฉgรฉnรฉrรฉe entre les fonctions analytiques sur $`U`$ et les distributions de support $`\{0\}`$. Plus abstraitement on peut remplacer les fonctions analytiques par les jets dโordre infini au point $`0`$. On appelle variรฉtรฉ formelle, ou voisinage formel de $`0`$, lโespace $`\overline{๐}`$ des distributions de support $`\{0\}`$. La structure dโalgรจbre commutative sur lโespace des fonctions analytiques sur $`U`$ dรฉtermine une structure de cogรจbre cocommutative sur son dual restreint, qui est exactement $`\overline{๐}`$. La comultiplication est donnรฉe par :
$$<\mathrm{\Delta }v,\phi \psi >=<v,\phi \psi >.$$
Considรฉrant lโespace tangent $`V`$ ร la variรฉtรฉ $`U`$ en $`0`$, on a en fait un isomorphisme de cogรจbres entre $`\overline{๐}`$ et $`S(V)`$, oรน la comultiplication $`\mathrm{\Delta }`$ de $`S(V)`$ est le morphisme dโalgรจbres tel que $`\mathrm{\Delta }(v)=v1+1v`$ pour $`vV`$. On considรจrera la version pointรฉe :
$$๐=S^+(V)=\underset{n1}{}S^n(V)$$
Cโest la cogรจbre colibre cocommutative sans co-unitรฉ construite sur $`V`$. Cโest aussi le dual restreint de lโalgรจbre des jets dโordre infini qui sโannulent en $`0`$. On remarque que $`\mathrm{\Delta }v=0`$ si et seulement si $`v`$ appartient ร $`V`$. Un champ de vecteurs sur la variรฉtรฉ formelle pointรฉe est donnรฉ par une codรฉrivation $`Q:๐๐`$ (cโest donc un champ de vecteurs qui sโannule en $`0`$). Un morphisme de variรฉtรฉs formelles pointรฉes est donnรฉ par un morphisme de cogรจbres. Tout morphisme $`f`$ de variรฉtรฉs pointรฉes induit un morphisme de variรฉtรฉs formelles par transport des distributions de support $`\{0\}`$ :
$$<f_{}T,\phi >=<T,\phi f>$$
Or, par propriรฉtรฉ universelle des cogรจbres cocommutatives colibres, une codรฉrivation $`Q:S^+(V)S^+(V)`$ (resp. un morphisme de cogรจbres $`:S^+(V_1)S^+(V_2)`$) est entiรจrement dรฉterminรฉ(e) par sa composition avec la projection sur $`V`$ (resp. $`V_2`$), cโest ร dire par une suite dโapplications :
$$Q_n:S^nVV\text{ (resp. }_n:S^nV_1V_2\text{)}$$
qui sont par dรฉfinition les coefficients de Taylor du champ de vecteurs $`Q`$ ou du morphisme $``$. III.2. Variรฉtรฉs formelles graduรฉes pointรฉes On fait la mรชme construction algรฉbrique dans la catรฉgorie des espaces vectoriels graduรฉs : une variรฉtรฉ formelle graduรฉe pointรฉe est une cogรจbre $`๐`$ isomorphe ร $`S^+(V)`$$`V`$ est cette fois-ci un espace graduรฉ. Toutes les notions du ยง III.1 sโappliquent, ร ceci prรจs que lโon peut considรฉrer des champs de vecteurs de diffรฉrents degrรฉs. Nous allons donner une formule explicite pour un champ de vecteurs ou un morphisme en fonction de ses coefficients de Taylor :
Thรฉorรจme III.2.1.
Soit $`i`$ un entier, soient $`V,V_1,V_2`$ des espaces graduรฉs, et deux suites dโapplications linรฉaires $`Q_n:S^nVV`$ de degrรฉ $`i`$, $`_n:S^nV_1V_2`$ de degrรฉ zรฉro. Alors il existe une unique codรฉrivation $`Q`$ de degrรฉ $`i`$ de $`S^+(V)`$ et un unique morphisme $`:S^+(V_1)S^+(V_2)`$ dont les $`Q_n`$ et les $`_n`$ sont les coefficients de Taylor respectifs. $`Q`$ et $``$ sont donnรฉs par les formules explicites :
$$\begin{array}{cc}\hfill Q(x_1\mathrm{}x_n)& =\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _x(I,J)\left(Q_{|I|}(x_I)\right).x_J\hfill \\ \hfill (x_1\mathrm{}x_n)& =\underset{j1}{}\frac{1}{j!}\underset{\genfrac{}{}{0pt}{}{I_1\mathrm{}I_j=\{1,\mathrm{},n\}}{I_1,\mathrm{},I_j\mathrm{}}}{}\epsilon _x(I_1,\mathrm{},I_j)_{|I_1|}(x_{I_1})\mathrm{}_{|I_j|}(x_{I_j})\hfill \end{array}$$
$`\epsilon _x(I_1,\mathrm{},I_j)`$ dรฉsigne la signature de lโeffet sur les $`x_i`$ impairs de la permutation-battement associรฉe ร la partition $`(I_1,\mathrm{},I_j)`$ de $`\{1,\mathrm{},n\}`$.
Dรฉmonstration. Supposons que tous les coefficients de Taylor de la codรฉrivation $`Q`$ sont nuls. En particulier $`Q(x)=0`$ pour tout $`xV`$. Supposons que $`Q(x_1\mathrm{}x_k)=0`$ pour tout $`kn`$ Alors :
$$\mathrm{\Delta }Q(x_1\mathrm{}x_{n+1})=(QI+IQ)\mathrm{\Delta }(x_1\mathrm{}x_{n+1})=0,$$
compte tenu de lโhypothรจse de rรฉcurrence et de lโexpression explicite de $`\mathrm{\Delta }(x_1\mathrm{}x_{n+1})`$ :
$$\mathrm{\Delta }(x_1\mathrm{}x_{n+1})=\underset{IJ=\{1,\mathrm{},n+1\},I,J\mathrm{}}{}\epsilon _x(I,J)x_Ix_J$$
Donc $`Q(x_1\mathrm{}x_{n+1})V`$, donc est nul puisque le $`n+1`$-รจme coefficient de Taylor est nul. Le raisonnement est analogue dans le cas dโun morphisme, et montre quโune codรฉrivation ou un morphisme est entiรจrement dรฉterminรฉ(e) par ses coefficients de Taylor. Nous vรฉrifions directement les formules (les vรฉrifications ร lโordre 2 ou 3 sont laissรฉes au lecteur ร titre dโexercice). 1. Cas dโune codรฉrivation : On รฉcrit la formule explicite pour $`\mathrm{\Delta }(x_1\mathrm{}x_n)`$ en utilisant la cocommutativitรฉ graduรฉe, ce qui permet de ne retenir que la moitiรฉ des partitions :
$$\mathrm{\Delta }(x_1\mathrm{}x_n)=(1+\tau )\underset{KL=\{1,\mathrm{},n\},1K,L\mathrm{}}{}\epsilon _x(K,L)x_Kx_L$$
On a donc, en prenant pour $`Q`$ lโexpression explicite du thรฉorรจme :
$$\begin{array}{cc}\hfill \mathrm{\Delta }Q(x_1& \mathrm{}x_n)=_{IJ=\{1,\mathrm{},n\},I,J\mathrm{}}\epsilon _x(I,J)\mathrm{\Delta }(Q_{|I|}x_I.x_J)\hfill \\ & =(1+\tau )\underset{IJ=\{1,\mathrm{},n\},I,J\mathrm{}}{}\underset{KL=J,L\mathrm{}}{}\epsilon _x(I,J)\epsilon _{x_J}(K,L)Q_{|I|}(x_I).x_Kx_L\hfill \\ & =(1+\tau )\underset{IJK=\{1,\mathrm{},n\},I,K\mathrm{}}{}\epsilon _x(I,J,K)Q_{|I|}(x_I).x_Jx_K\hfill \end{array}$$
Par ailleurs on a :
$$\begin{array}{cc}\hfill (Q& I+IQ)\mathrm{\Delta }(x_1\mathrm{}x_n)=(1+\tau )_{LK=\{1,\mathrm{},n\},L,K\mathrm{}}\epsilon _x(L,K)Q(x_L)x_K\hfill \\ & =(1+\tau )\underset{LK=\{1,\mathrm{},n\},L,K\mathrm{}}{}\epsilon _x(L,K)\underset{IJ=L,I,J\mathrm{}}{}\epsilon _{x_L}(I,J)Q_{|I|}(x_I).x_Jx_K\hfill \\ & =(1+\tau )\underset{IJK=\{1,\mathrm{},n\},,I,J,K\mathrm{}}{}\epsilon _x(I,J,K)Q_{|I|}(x_I).x_Jx_K\hfill \end{array}$$
dโoรน le fait que $`Q`$ est bien une codรฉrivation. 2. Cas dโun morphisme : le calcul est un peu plus compliquรฉ : on commence par รฉcrire $`\mathrm{\Delta }`$ et $``$ de maniรจre redondante, en employant des permutations qui ne sont pas forcรฉment des battements :
$$\begin{array}{cc}\hfill \mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{\sigma S_n}{}\underset{r=1}{\overset{n}{}}\frac{\epsilon _x(\sigma )}{r!(nr)!}x_{\sigma _1}\mathrm{}x_{\sigma _r}x_{\sigma _{r+1}}\mathrm{}x_{\sigma _n}\hfill \\ \hfill (x_1\mathrm{}x_n)& =\underset{j1}{}\frac{1}{j!}\underset{k_1+\mathrm{}+k_j=n}{}\frac{1}{k_1!\mathrm{}k_j!}\underset{\sigma S_n}{}\epsilon _x(\sigma )_{k_1}(x_{\sigma _1}\mathrm{}x_{\sigma _{k_1}})\mathrm{}_{k_j}(x_{\mathrm{}}\mathrm{}x_{\sigma _n})\hfill \end{array}$$
On vรฉrifie directement lโรฉgalitรฉ :
$$\mathrm{\Delta }(x_1\mathrm{}x_n)=()\mathrm{\Delta }(x_1\mathrm{}x_n)$$
Lโรฉcriture par blocs : $`x_1\mathrm{}x_n=(x_1\mathrm{}x_{k_1})\mathrm{}(x_{k_1+\mathrm{}+k_{j1}+1}\mathrm{}x_n)`$ induit par permutation des blocs un plongement du groupe de permutations $`S_j`$ dans $`S_n`$. On calcule :
$$\begin{array}{cc}\hfill \mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{j2}{}\frac{1}{j!}\underset{k_1+\mathrm{}+k_j=n}{}\frac{1}{k_1!\mathrm{}k_j!}\underset{\sigma S_n}{}\epsilon _x(\sigma )\mathrm{\Delta }\left(_{k_1}(x_{\sigma _1}\mathrm{}x_{\sigma _{k_1}})\mathrm{}_{k_j}(x_{\mathrm{}}\mathrm{}x_{\sigma _n})\right)\hfill \\ & =\underset{j2}{}\frac{1}{j!}\underset{k_1+\mathrm{}+k_j=n}{}\frac{1}{k_1!\mathrm{}k_j!}\underset{\sigma S_n}{}\epsilon _x(\sigma )\underset{\tau S_jS_n}{}\hfill \\ & \underset{r=1}{\overset{j1}{}}\frac{\epsilon _{\sigma x}(\tau )}{r!(jr)!}_{k_{\tau _1}}(\mathrm{})\mathrm{}_{k_{\tau _r}}(\mathrm{})_{k_{\tau _{r+1}}}(\mathrm{})\mathrm{}_{k_{\tau _j}}(\mathrm{})\hfill \end{array}$$
Dans le dernier membre de lโรฉgalitรฉ ci-dessus, chaque terme se trouve rรฉpรฉtรฉ autant de fois quโil y a dโรฉlรฉments dans $`S_j`$. On a donc :
$$\begin{array}{cc}\hfill \mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{j2}{}\underset{k_1+\mathrm{}+k_j=n}{}\frac{1}{k_1!\mathrm{}k_j!}\underset{\sigma S_n}{}\epsilon _x(\sigma )\hfill \\ & \underset{r=1}{\overset{j1}{}}\frac{1}{r!(jr)!}_{k_1}(\mathrm{})\mathrm{}_{k_r}(\mathrm{})_{k_{r+1}}(\mathrm{})\mathrm{}_{k_j}(\mathrm{})\hfill \end{array}$$
Par ailleurs, on a :
$$\begin{array}{cc}\hfill ()\mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{\sigma S_n}{}\underset{r=1}{\overset{n1}{}}\frac{\epsilon _x(\sigma )}{r!(nr)!}(x_{\sigma _1}\mathrm{}x_{\sigma _r})(x_{\sigma _{r+1}}\mathrm{}x_{\sigma _n})\hfill \\ & =\underset{\sigma S_n}{}\underset{r=1}{\overset{n1}{}}\underset{\alpha S_r\times S_{nr}S_n}{}\frac{\epsilon _x(\sigma )\epsilon _{\sigma x}(\alpha )}{r!(nr)!}\hfill \\ & \underset{j,k1}{}\frac{1}{j!k!}\underset{\genfrac{}{}{0pt}{}{r_1+\mathrm{}+r_j=r}{s_1+\mathrm{}+s_k=nr}}{}\frac{1}{r_1!\mathrm{}r_j!s_1!\mathrm{}s_k!}\hfill \\ & _{r_1}(x_{\alpha _{\sigma _1}}\mathrm{}x_{\alpha _{\sigma _{r1}}})\mathrm{}_{r_j}(\mathrm{})_{s_1}(\mathrm{})\mathrm{}_{s_k}(\mathrm{}x_{\alpha _{\sigma _n}})\hfill \end{array}$$
Dans le dernier membre de lโรฉgalitรฉ ci-dessus, chaque terme se trouve rรฉpรฉtรฉ autant de fois quโil y a dโรฉlรฉments dans $`S_r\times S_{nr}`$. Donc :
$$\begin{array}{cc}\hfill ()\mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{\sigma S_n}{}\underset{r=1}{\overset{n1}{}}\epsilon _x(\sigma )\frac{1}{j!k!}\underset{\genfrac{}{}{0pt}{}{r_1+\mathrm{}+r_j=r}{s_1+\mathrm{}+s_k=nr}}{}\frac{1}{r_1!\mathrm{}r_j!s_1!\mathrm{}s_k!}\hfill \\ & _{r_1}(x_{\sigma _1}\mathrm{}x_{\sigma _{r_1}})\mathrm{}_{r_j}(\mathrm{})_{s_1}(\mathrm{})\mathrm{}_{s_k}(\mathrm{}x_{\sigma _n})\hfill \end{array}$$
Posant $`l=j+k`$ et procรฉdant ร la renumรฉrotation $`(s_1,\mathrm{}s_k)=(r_{j+1},\mathrm{}r_l)`$ on obtient :
$$\begin{array}{cc}\hfill ()\mathrm{\Delta }(x_1\mathrm{}x_n)& =\underset{\sigma S_n}{}\epsilon _x(\sigma )\underset{l2}{}\underset{r_1+\mathrm{}+r_l=n}{}\underset{k=1}{\overset{l1}{}}\frac{1}{k!(lk)!}\frac{1}{r_1!\mathrm{}r_l!}\hfill \\ & _{r_1}(x_{\sigma _1}\mathrm{}x_{\sigma _{r1}})\mathrm{}_{r_k}(\mathrm{})_{r_{k+1}}(\mathrm{})\mathrm{}_{r_l}(\mathrm{}x_{\sigma _n})\hfill \\ & =\mathrm{\Delta }(x_1\mathrm{}x_n)\hfill \end{array}$$
compte tenu du calcul prรฉcรฉdent, ce qui dรฉmontre le thรฉorรจme.
$``$
IV. L-algรจbres et L-morphismes A tout espace vectoriel graduรฉ $`V`$ on associe (attention au dรฉcalage!) la variรฉtรฉ formelle $`(V[1],0)`$ pointรฉe, cโest ร dire la cogรจbre colibre sans co-unitรฉ :
$$๐(V)=S^+(V[1])\stackrel{~}{\underset{\mathrm{\Phi }}{}}\underset{k1}{}(\mathrm{\Lambda }^kV)[k]$$
$`\mathrm{\Phi }`$ est lโisomorphisme dรฉcrit au ยง II.4. Un prรฉ-$`L_{\mathrm{}}`$-morphisme entre deux espaces graduรฉs $`V_1`$ et $`V_2`$ est par dรฉfinition un morphisme de variรฉtรฉs formelles, cโest-ร -dire un morphisme de cogรจbres :
$$:๐(V_1)๐(V_2)$$
qui est donc dรฉterminรฉ par ses coefficients de Taylor $`_j`$. Posant $`\overline{}_j=_j\mathrm{\Phi }^1`$ on a :
$$\begin{array}{c}\text{ }\overline{}_1:V_1V_2\text{ }\hfill \\ \text{ }\overline{}_2:\mathrm{\Lambda }^2V_1V_2[1]\text{ }\hfill \\ \text{ }\overline{}_3:\mathrm{\Lambda }^3V_1V_2[2]\text{ }\hfill \\ \text{ }\mathrm{}\text{ }\hfill \end{array}$$
IV.1. Algรจbres de Lie homotopiques Par dรฉfinition une $`L_{\mathrm{}}`$-algรจbre, ou algรจbre de Lie homotopique est une variรฉtรฉ formelle graduรฉe pointรฉe du type $`(\text{g}[1],0)`$, oรน g est un espace vectoriel graduรฉ, munie dโun champ de vecteurs $`Q`$ de degrรฉ $`1`$ vรฉrifiant lโรฉquation maรฎtresse :
$$[Q,Q]=2Q^2=0$$
Cโest-ร -dire que $`Q`$ est une codรฉrivation de carrรฉ nul de la cogรจbre $`๐(\text{g})`$. Les coefficients de Taylor $`Q_k:S^k(\text{g}[1])\text{g}[2]`$ donnent naissance aux coefficients $`\overline{Q}_k=Q_k\mathrm{\Phi }^1`$ :
$$\begin{array}{c}\text{ }\overline{Q}_1:\text{g}\text{g}[1]\text{ }\hfill \\ \text{ }\overline{Q}_2:\mathrm{\Lambda }^2\text{g}\text{g}\text{ }\hfill \\ \text{ }\overline{Q}_3:\mathrm{\Lambda }^3\text{g}\text{g}[1]\text{ }\hfill \\ \text{ }\mathrm{}\text{ }\hfill \end{array}$$
Lโรฉquation maรฎtresse se traduit par une infinitรฉ de relations quadratiques entre les $`\overline{Q}_k`$, qui sโobtiennent en รฉcrivant explicitement pour tout $`k`$ lโรฉquation :
$$\pi Q^2(x_1\mathrm{}x_k)=0$$
$`\pi :๐(\text{g})\text{g}[1]`$ est la projection canonique. On รฉcrit explicitement les trois premiรจres : Premiรจre รฉquation : $`Q_1^2(x)=0`$ pour tout $`x`$ dans g. Donc $`(\text{g},Q_1)`$ est un complexe de cochaรฎnes. Deuxiรจme รฉquation : $`\pi Q^2(x.y)=0`$, soit :
$$Q_2(Q_1x.y+(1)^{|x|1}x.Q_1y)+Q_1Q_2(x.y)=0.$$
(Remarque : $`|x|1`$ est bien le degrรฉ de $`x`$ dans la cogรจbre $`๐(\text{g})`$, ร cause du dรฉcalage). Traduisant cette รฉgalitรฉ en termes de $`\overline{Q}_1`$ et $`\overline{Q}_2`$ on obtient (cf ยง II.4) :
$$\alpha (\overline{Q}_1x,y)\overline{Q}_2(\overline{Q}_1xy)+(1)^{|x|1}\alpha (x,\overline{Q}_1y)\overline{Q}_2(x\overline{Q}_1y)+\alpha (x,y)\overline{Q}_1\overline{Q}_2(xy)=0$$
Compte tenu de lโรฉgalitรฉ :
$$\alpha (x,y)=(1)^{|x|(|y|1)}$$
on obtient :
$$(1)^{|y|1}\overline{Q}_2(\overline{Q}_1xy)\overline{Q}_2(x\overline{Q}_1y)+\overline{Q}_1\overline{Q}_2(xy)=0.$$
Posant $`dx=(1)^{|x|}\overline{Q}_1x`$ et $`[x,y]=\overline{Q}_2(xy)`$ on obtient finalement :
$$d[x,y]=[dx,y]+(1)^{|x|}[x,dy]$$
donc $`\overline{Q}_2`$ est un crochet antisymรฉtrique pour lequel $`d`$ est une dรฉrivation. Remarque : On peut garder $`\overline{Q}_1`$ comme dรฉrivation sans le modifier, ร condition dโinverser le sens du crochet, cโest-ร -dire de poser :
$$[x,y]=\overline{Q}_2(yx).$$
Nous choisirons la premiรจre solution. Troisiรจme รฉquation : $`\pi Q_3(x.y.z)=0`$ soit :
$$\begin{array}{c}\text{ }Q_3\left(Q_1x.y.z+(1)^{|x|1}x.Q_1y.z+(1)^{|x|+|y|2}x.y.Q_1z\right)+Q_1Q_3(x.y.z)\text{ }\hfill \\ \text{ }+Q_2\left(Q_2(x.y).z+(1)^{(|y|1)(|z|1)}Q_2(x.z).y+(1)^{(|x|1)(|y|+|z|2)}Q_2(y.z).x\right)=0\text{ }\hfill \end{array}$$
soit :
$$\begin{array}{c}\text{ }Q_2(Q_2(x.y).z+(1)^{(|y|1)(|z|1)+(|x|1)(|z|1)}Q_2(z.x).y\text{ }\hfill \\ \text{ }+(1)^{(|x|1)(|y|+|z|)}Q_2(y.z).x)+\text{ }\text{termes en }Q_3\text{ }=0\text{ }\hfill \end{array}$$
Or on a :
$$\begin{array}{cc}\hfill Q_2(Q_2(x.y).z)& =\alpha (Q_2(x.y),z)\alpha (x,y)\overline{Q}_2(\overline{Q}_2(xy)z)\hfill \\ & =(1)^{(|x|+|y|)|z|}(1)^{(|x|1)|y|}\overline{Q}_2\left(\overline{Q}_2(xy)z\right)\hfill \end{array}$$
En reportant ceci dans lโรฉquation prรฉcรฉdente et en simplifiant par $`(1)^{|x||y|+|x||z|+|y||z|}`$ on obtient finalement :
$$(1)^{|x||z|}[[x,y],z]+(1)^{|y||x|}[[y,z],x]+(1)^{|z||y|}[[z,x],y]+\text{ }\text{termes en }Q_3\text{ }=0$$
Autrement dit le crochet fourni par $`\overline{Q}_2`$ vรฉrifie lโidentitรฉ de Jacobi graduรฉe โร homotopie gouvernรฉe par $`\overline{Q}_3`$ prรจsโ. En corollaire :
Thรฉorรจme IV.1.1.
Une algรจbre de Lie diffรฉrentielle graduรฉe est la mรชme chose quโune $`L_{\mathrm{}}`$-algรจbre pour laquelle tous les coefficients de Taylor sont nuls sauf les deux premiers.
IV.2. Lโalgรจbre de Lie diffรฉrentielle graduรฉe des mutichamps de vecteurs Sur $`V=T_{poly}\left(\text{}^d\right)`$, on dispose du crochet de Schouten dรฉfini par :
$$[\xi _1\mathrm{}\xi _k,\eta _1\mathrm{}\eta _{\mathrm{}}]_S=\underset{i=1}{\overset{k}{}}\underset{j=1}{\overset{\mathrm{}}{}}(1)^{i+j}[\xi _i,\eta _j]\xi _1\mathrm{}\widehat{\xi _i}\mathrm{}\xi _k\eta _1\mathrm{}\widehat{\eta _j}\mathrm{}\eta _{\mathrm{}}.$$
La symรฉtrisation de $`Tens\left(\text{}^d\right)`$ nous permet de dรฉfinir une opรฉration $``$.
Si $`\alpha _1`$ est un $`k_1`$-tenseur antisymรฉtrique :
$$\alpha _1=\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1}_{i_2}\mathrm{}_{i_{k_1}}Tens^{k_1}\left(\text{}^d\right),$$
alors :
$$\mathrm{\Phi }_{k_1}^1(\alpha _1)=\alpha _1^{i_1\mathrm{}i_{k_1}}\psi _{i_1}\mathrm{}\psi _{i_{k_1}}S\left(๐ณ\left(\text{}^d\right)[1]\right)[k_1]$$
oรน chaque $`\psi _i=\mathrm{\Phi }_1^1(_i)`$ est une variable de degrรฉ 1.
Si maintenant $`\alpha _2`$ est un $`k_2`$ tenseur antisymรฉtrique, on posera :
$$\alpha _1\alpha _2=\mathrm{\Phi }_{k_1+k_21}\left(\underset{i=1}{\overset{d}{}}\frac{\mathrm{\Phi }_{k_1}^1(\alpha _1)}{\psi _i}\text{.}\frac{\mathrm{\Phi }_{k_2}^1(\alpha _2)}{x_i}\right)$$
en tenant compte du fait que $`\frac{}{\psi _i}`$ est un opรฉrateur de dรฉrivation impair.
Lemme IV.2.1 (Calcul de $`\alpha _1\alpha _2`$).
On a :
$$\alpha _1\alpha _2=\underset{l=1}{\overset{k_1}{}}(1)^{l1}\alpha _1^{i_1\mathrm{}.i_{k_1}}_l\alpha _2^{j_1\mathrm{}j_{k_2}}_{i_1}\mathrm{}\widehat{_{i_l}}\mathrm{}_{i_{k_1}}_{j_1}\mathrm{}_{j_{k_2}}$$
et
$$[\alpha _1,\alpha _2]_S=(1)^{k_11}\alpha _1\alpha _2(1)^{k_1(k_21)}\alpha _2\alpha _1.$$
Dรฉmonstration. On a :
$$\begin{array}{cc}\hfill \frac{}{\psi _i}\left(\alpha _1^{i_1\mathrm{}i_{k_1}}\psi _{i_1}\text{.}\mathrm{}\text{.}\psi _{i_{k_1}}\right)& =\underset{l=1}{\overset{k_1}{}}(1)^{l1}\alpha _1^{i_1\mathrm{}i_{k_1}}\psi _{i_1}\text{.}\mathrm{}\text{.}\frac{\psi _{i_l}}{\psi _i}\text{.}\mathrm{}\text{.}\psi _{i_{k_1}}\hfill \\ & =\underset{l=1}{\overset{k_1}{}}(1)^{l1}\delta _{i_l}^i\alpha _1^{i_1\mathrm{}i_{k_1}}\psi _{i_1}\text{.}\mathrm{}\text{.}\widehat{\psi _{i_l}}\text{.}\mathrm{}\text{.}\psi _{i_{k_1}}.\hfill \end{array}$$
Donc :
$$\begin{array}{cc}\hfill \underset{i=1}{\overset{d}{}}\frac{\mathrm{\Phi }_{k_1}^1(\alpha _1)}{\psi _i}\text{.}\frac{\mathrm{\Phi }_{k_2}^1(\alpha _2)}{x_i}& =\underset{l=1}{\overset{k_1}{}}(1)^{l1}\alpha _1^{i_1\mathrm{}i_{k_1}}\psi _{i_1}\text{.}\mathrm{}\text{.}\widehat{\psi _{i_l}}\text{.}\mathrm{}\text{.}\psi _{i_{k_1}}\text{.}_{i_l}\alpha _2^{j_1\mathrm{}j_{k_2}}\psi _{j_1}\text{.}\mathrm{}\text{.}\psi _{j_{k_2}}\hfill \\ & =\underset{l=1}{\overset{k_1}{}}(1)^{l1}\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_l}\alpha _2^{j_1\mathrm{}j_{k_2}}\psi _{i_1}\text{.}\mathrm{}\text{.}\widehat{\psi _{i_l}}\text{.}\mathrm{}\text{.}\psi _{i_{k_1}}\text{.}\psi _{j_1}\text{.}\mathrm{}\text{.}\psi _{j_{k_2}}.\hfill \end{array}$$
Dโautre part :
$$\begin{array}{cc}\hfill [\alpha _1,\alpha _2]_S& =[\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1}\mathrm{}_{i_{k_1}},\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_1}\mathrm{}_{j_{k_2}}]\hfill \\ & =[\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1},\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_1}]_{i_2}\mathrm{}_{i_{k_1}}_{j_1}\mathrm{}_{j_{k_2}}+\hfill \\ & +\underset{l=2}{\overset{k_2}{}}(1)^{1+l}[\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1},_{j_l}]_{i_2}\mathrm{}_{i_{k_1}}\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_1}\mathrm{}\widehat{_{j_l}}\mathrm{}_{j_{k_2}}+\hfill \\ & +\underset{l=2}{\overset{k_1}{}}[_{i_l},\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_1}]\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1}\mathrm{}\widehat{_{i_l}}\mathrm{}_{i_{k_1}}_{j_2}\mathrm{}_{j_{k_2}}\hfill \\ & =\underset{l=1}{\overset{k_2}{}}(1)^{l+1}\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_l}\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_1}\mathrm{}_{i_{k_1}}_{j_1}\mathrm{}\widehat{_{j_l}}\mathrm{}_{j_{k_2}}+\hfill \\ & +\underset{l=1}{\overset{k_1}{}}(1)^{l+1}\alpha _1^{i_1\mathrm{}i_{k_1}}_{i_l}\alpha _2^{j_1\mathrm{}j_{k_2}}_{j_1}_{i_1}\mathrm{}\widehat{_{i_l}}\mathrm{}_{i_{k_1}}_{j_2}\mathrm{}\mathrm{}_{j_{k_2}}\hfill \\ & =(1)^{k_11}\alpha _1\alpha _2(1)^{(k_21)k_1}\alpha _2\alpha _1.\hfill \end{array}$$
Corollaire IV.2.2.
Lโespace graduรฉ $`T_{poly}\left(\text{}^d\right)`$, muni du crochet :
$$[\alpha _1,\alpha _2]_S^{}=[\alpha _2,\alpha _1]_S$$
est aussi une algรจbre de Lie graduรฉe et :
$$[\alpha _1,\alpha _2]_S^{}=(1)^{(k_11)k_2}\alpha _1\alpha _2+(1)^{k_2}\alpha _2\alpha _1.$$
Comme $`[,]_S^{}`$ dรฉfinit sur $`T_{poly}\left(\text{}^d\right)`$ une structure dโalgรจbre de Lie graduรฉe, on aura, en prenant $`d=0`$, une structure de $`L_{\mathrm{}}`$ algรจbre sur $`๐\left(T_{poly}\left(\text{}^d\right)\right)`$. Le champ de vecteurs $`Q`$ est caractรฉrisรฉ par :
$$\begin{array}{cc}\hfill Q_1=0,Q_2(\alpha _1\text{.}\alpha _2)& =(1)^{(k_11)k_2}[\alpha _1,\alpha _2]_S^{}\hfill \\ & =\alpha _1\alpha _2+(1)^{k_1k_2}\alpha _2\alpha _1.\hfill \end{array}$$
IV.3. Lโalgรจbre de Lie diffรฉrentielle graduรฉe des opรฉrateurs polydiffรฉrentiels On considรจre lโespace vectoriel $`V^{}=D_{poly}\left(\text{}^d\right)`$ des (combinaisons linรฉaires dโ) opรฉrateurs multidiffรฉrentiels graduรฉ par $`|A|=m1`$ si $`A`$ est $`m`$-diffรฉrentiel.
Sur $`D_{poly}\left(\text{}^d\right)`$, lโopรฉrateur de composition naturel $``$ sโรฉcrit :
$$\begin{array}{cc}\hfill \left(A_1A_2\right)& (f_1,\mathrm{},f_{m_1+m_21})=\hfill \\ & =\underset{j=1}{\overset{m_1}{}}(1)^{(m_21)(j1)}A_1(f_1,\mathrm{},f_{j1},A_2(f_j,\mathrm{},f_{j+m_21}),f_{j+m_2},\mathrm{},f_{m1+m_21}).\hfill \end{array}$$
On associe ร cette composition dโune part le crochet de Gerstenhaber :
$$[A_1,A_2]_G=A_1A_2(1)^{|A_1||A_2|}A_2A_1,$$
dโautre part lโopรฉrateur de cobord :
$$dA=[\mu ,A]$$
$`\mu `$ est la multiplication des fonctions : $`\mu (f_1,f_2)=f_1f_2`$. Remarque : Avec ce choix de $`d`$, $`(D_{poly}\left(\text{}^d\right),[,]_G,d)`$ est une algรจbre de Lie graduรฉe diffรฉrentielle, on vรฉrifie en effet que $`dd=0`$ et
$$d\left([A_1,A_2]\right)=[dA_1,A_2]+(1)^{|A_1|}[A_1,dA_2].$$
Lโopรฉrateur de cobord de Hochschild usuel $`d_H`$ donnรฉ par :
$$\begin{array}{cc}\hfill \left(d_HA\right)(f_1,\mathrm{},f_m)& =f_1A(f_2,\mathrm{},f_m)A(f_1f_2,f_3,\mathrm{},f_m)+\mathrm{}+(1)^mA(f_1,\mathrm{},f_{m1})f_m\hfill \\ & =(1)^{|A|+1}dA(f_1,\mathrm{},f_m)\hfill \end{array}$$
nโest pas une dรฉrivation de lโalgรจbre de Lie graduรฉe $`(D_{poly}\left(\text{}^d\right),[,]_G)`$.
Le champ de vecteurs $`Q^{}`$ sur la variรฉtรฉ formelle $`๐(V^{})`$ sera donc dรฉfini par :
$$Q_1^{}(A)=(1)^{|A|}dA=(1)^{|A|+1}[\mu ,A]=[A,\mu ]=d_HA$$
et
$$\begin{array}{cc}\hfill Q_2^{}\left(A_1\text{.}A_2\right)& =(1)^{|A_1|(|A_2|1)}[A_1,A_2]_G\hfill \\ & =(1)^{|A_1|(|A_2|1)}A_1A_2(1)^{|A_1|}A_2A_1.\hfill \end{array}$$
IV.4. L-morphismes Par dรฉfinition un $`L_{\mathrm{}}`$-morphisme entre deux $`L_{\mathrm{}}`$-algรจbres $`(\text{g}_1,Q)`$ et $`(\text{g}_2,Q^{})`$ est un morphisme de variรฉtรฉs formelles pointรฉes :
$$:๐(\text{g}_1)๐(\text{g}_2)$$
vรฉrifiant :
$$Q=Q^{}$$
Cette รฉquation induit une infinitรฉ de relations entre les coefficients de Taylor de $`Q`$, $`Q^{}`$ et $``$, dont nous allons examiner les deux premiรจres : Premiรจre รฉquation : $`Q_1^{}_1(x)=_1Q_1(x)`$, cโest-ร -dire que $`_1`$ est un morphisme de complexes. Deuxiรจme รฉquation : $`\pi Q^{}(x.y)=\pi Q(x.y)`$ soit :
$$\pi Q^{}(_1x._1y+_2(x.y))=\pi (Q_1x.y+(1)^{(|x|1)}x.Q_1y+Q_2(x.y))$$
soit encore :
$$Q_2^{}(_1x._1y)+Q_1^{}_2(x.y)=_2(Q_1x.y+(1)^{|x|1}x.Q_1y)+_1Q_2(x.y).$$
On traduit cette derniรจre รฉgalitรฉ en termes de $`\overline{Q}_1`$, $`\overline{Q}_2`$, $`\overline{}_1`$, etc. :
$$\begin{array}{c}\text{ }(1)^{|x|(|y|1)}[\overline{}_1x,\overline{}_1y]+(1)^{|x|+|y|1+|x|(|y|1)}d\overline{}_2(xy)\text{ }\hfill \\ \text{ }=(1)^{(|x|1)(|y|1)+|x|}\overline{}_2(dxy)+(1)^{|x||y|+|x|1+|y|}\overline{}_2(xdy)+(1)^{|x|(|y|1)}\overline{}_1([x,y])\text{ }\hfill \end{array}$$
soit, en multipliant par $`(1)^{|x|(|y|1)}`$ :
$$\overline{}_1([x,y])[\overline{}_1x,\overline{}_1y]=(1)^{|x|+|y|1}\left(d\overline{}_2(xy)\overline{}_2(dxy)(1)^{|x|}\overline{}_2(xdy)\right).$$
Dans le cas oรน $`\text{g}_1`$ et $`\text{g}_2`$ sont des algรจbres de Lie diffรฉrentielles graduรฉes, $`_1`$ nโest donc pas forcรฉment un morphisme dโalgรจbres de lie diffรฉrentielles graduรฉes, mais le dรฉfaut est gouvernรฉ par le coefficient suivant, cโest-ร -dire $`_2`$.
Proposition IV.4.1 (Equation de $`L_{\mathrm{}}`$-morphisme dans le cas des algรจbres de Lie diffรฉrentielles graduรฉes).
Supposons que $`(V,[,],d)`$ et $`(V^{},[,]^{},d^{})`$ soient deux algรจbres de Lie graduรฉes. Notons $`(๐(V),Q)`$ et $`(๐(V^{}),Q^{})`$ les $`L_{\mathrm{}}`$ algรจbres correspondantes respectives. Soit $`:๐(V)๐(V^{})`$ un morphisme de cogรจbre. Alors $``$ est un $`L_{\mathrm{}}`$ morphisme si et seulement si :
$$\begin{array}{cc}\hfill Q_1^{}_n\left(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n\right)+& \frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (I,J)Q_2^{}\left(_{|I|}\left(\alpha _I\right)\text{.}_{|J|}\left(\alpha _J\right)\right)=\hfill \\ & =\underset{k=1}{\overset{n}{}}\epsilon _\alpha (k,1,\mathrm{}\widehat{k},\mathrm{},n)_n\left(Q_1(\alpha _k)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _k}\text{.}\mathrm{}\text{.}\alpha _n\right)+\hfill \\ & +\frac{1}{2}\underset{kl}{}\epsilon _\alpha (k,l,1,\mathrm{},\widehat{k,l},\mathrm{},n)_{n1}\left(Q_2(\alpha _k\text{.}\alpha _l)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _k}\text{.}\mathrm{}\text{.}\widehat{\alpha _l}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \end{array}$$
$`|I|`$ et $`\epsilon _\alpha (I,J)`$ ont la mรชme signification que dans le thรฉorรจme III.2.1, et oรน $`\epsilon _\alpha (\mathrm{})`$ dรฉsigne le signe de Quillen de la permutation indiquรฉe entre parenthรจses, cโest-ร -dire la signature de la trace sur les $`\alpha _j`$ impairs de cette permutation.
Comme pour les codรฉrivations $`Q`$ et les morphismes de cogรจbres $``$, il est facile de voir que les applications $`Q^{}`$ et $`Q`$ sont uniquement dรฉterminรฉes par leur composition avec la projection sur $`V^{}[1]`$. On dรฉduit alors lโรฉquation de $`L_{\mathrm{}}`$-morphisme sous la forme $`(Q^{})_n=(Q)_n`$ pour tout $`n`$. Puisquโon est parti de deux algรจbres de Lie diffรฉrentielles graduรฉes, tous les $`Q_p`$ et $`Q_p^{}`$ sont nuls pour $`p3`$.
V. Quasi-isomorphismes Par dรฉfinition un quasi-isomorphisme entre deux $`L_{\mathrm{}}`$-algรจbres $`(\text{g}_1,Q_1)`$ et $`(\text{g}_2,Q_2)`$ est un $`L_{\mathrm{}}`$-morphisme $``$ dont le premier coefficient de Taylor $`_1:\text{g}_1[1]\text{g}_2[1]`$ est un morphisme de complexes qui induit un isomorphisme en cohomologie (quasi-isomorphisme de complexes). Nous allons exposer la dรฉmonstration du thรฉorรจme suivant (\[K1\] theorem 4.4) :
Thรฉorรจme V.1.
Pour tout quasi-isomorphisme $``$ dโune $`L_{\mathrm{}}`$-algรจbre $`(\text{g}_1,Q_1)`$ vers une $`L_{\mathrm{}}`$-algรจbre $`(\text{g}_2,Q_2)`$ il existe un $`L_{\mathrm{}}`$-morphisme $`๐ข`$ de $`(\text{g}_2,Q_2)`$ vers $`(\text{g}_1,Q_1)`$ dont le premier coefficient de Taylor $`๐ข_1:\text{g}_2[1]\text{g}_1[1]`$ soit un quasi-inverse pour $`_1`$.
V.1. Dรฉcomposition des $`L_{\mathrm{}}`$-algรจbres Une $`L_{\mathrm{}}`$-algรจbre $`(\text{g},Q)`$ est minimale si $`Q_1=0`$. Une $`L_{\mathrm{}}`$-algรจbre est linรฉaire contractile si $`Q_j=0`$ pour $`j2`$ et si la cohomologie du complexe donnรฉ par $`Q_1`$ est triviale. On remarque que la premiรจre notion est invariante par $`L_{\mathrm{}}`$-isomorphismes, contrairement ร la seconde notion.
Proposition V.2.
Toute $`L_{\mathrm{}}`$-algรจbre $`(\text{g},Q)`$ est $`L_{\mathrm{}}`$-isomorphe ร la somme directe dโune $`L_{\mathrm{}}`$-algรจbre minimale et dโune $`L_{\mathrm{}}`$-algรจbre linรฉaire contractile.
Dรฉmonstration. On dรฉcompose le complexe $`(\text{g},Q_1)`$ en somme directe $`(\text{g}^{},M_1)(\text{g}^{\prime \prime },L_1)`$$`M_1`$ est une diffรฉrentielle nulle et oรน $`(\text{g}^{\prime \prime },L_1)`$ est un complexe ร cohomologie triviale (on nรฉglige le dรฉcalage qui nโest pas essentiel ici). Pour ce faire on note comme dโhabitude $`Z_k`$ et $`B_k`$ le noyau et lโimage de la diffรฉrentielle en degrรฉ $`k`$, on choisit un supplรฉmentaire $`\text{g}_k^{}`$ de $`B_k`$ dans $`Z_k`$, et un supplรฉmentaire $`W_k`$ de $`Z_k`$ dans $`\text{g}_k`$. Posant alors $`\text{g}_k^{\prime \prime }=B_kW_k`$ on a la dรฉcomposition cherchรฉe. Cette dรฉcomposition du complexe est le point de dรฉpart de la dรฉcomposition de la $`l_{\mathrm{}}`$-algรจbre $`(\text{g},Q)`$. La cogรจbre associรฉe ร $`\text{g}=\text{g}^{}\text{g}^{\prime \prime }`$ sโรฉcrit :
$$๐(\text{g})=๐(\text{g}^{})๐(\text{g}^{\prime \prime })๐(\text{g}^{})๐(\text{g}^{\prime \prime }).$$
Il sโagit de construire un isomorphisme de cogรจbres :
$$:๐(\text{g})\stackrel{~}{}๐(\text{g})$$
tel que $`Q=\overline{Q}`$, avec :
$$\begin{array}{cc}& \overline{Q}_|๐\left(\text{g}^{}\right)=M\hfill \\ & \overline{Q}_|๐\left(\text{g}^{\prime \prime }\right)=L\hfill \\ & \overline{Q}_|๐\left(\text{g}^{}\right)๐\left(\text{g}^{\prime \prime }\right)=MI+IL\hfill \end{array}$$
$`M_1=0`$, $`L_j=0`$ pour $`j2`$ et $`L_1`$ ร cohomologie triviale. On pose donc pour commencer $`_1=Id:\text{g}\text{g}`$, dโoรน forcรฉment $`\overline{Q}_1=Q_1=L_1`$ au vu de la dรฉcomposition du complexe rappelรฉe ci-dessus. Il est trรจs facile de voir quโun $`L_{\mathrm{}}`$-morphisme $``$ vรฉrifiant $`_1=\text{Id}`$ sโรฉcrit comme un produit infini :
$$=\mathrm{}^k^{k1}\mathrm{}^2$$
$`^k`$ est le $`L_{\mathrm{}}`$-morphisme ayant lโidentitรฉ comme premier coefficient de Taylor, $`_k`$ comme $`k^{\text{iรจme}}`$ coefficient de Taylor, tous les autres coefficients รฉtant nuls. Chercher le coefficient $`_k`$ en supposant que les $`_j`$ sont connus pour $`j<k`$, cโest donc chercher un $`L_{\mathrm{}}`$-isomorphisme $``$ โlacunaireโ comme le $`^k`$ ci-dessus, entre $`(\text{g}^{}\text{g}^{\prime \prime },Q)`$ et $`(\text{g}^{}\text{g}^{\prime \prime },\overline{Q})`$, oรน le champ de vecteurs impair $`Q`$ vรฉrifie :
$$\begin{array}{cc}& Q_1{}_{|}{}^{}๐\left(\text{g}^{}\right)=0\hfill \\ & Q_j\left(๐(\text{g}^{})\right)\text{g}^{}\text{ pour }jk1\hfill \\ & Q_j\left(๐(\text{g}^{\prime \prime })\right)=0\text{ pour }2jk1\hfill \\ & Q_j\left(๐(\text{g}^{})๐(\text{g}^{\prime \prime })\right)=0\text{ pour }jk1\hfill \end{array}$$
et oรน le champ de vecteurs $`\overline{Q}`$ vรฉrifie les mรชmes conditions avec $`k`$ ร la place de $`k1`$. On supposera รฉgalement que les coefficients de Taylor de $`Q`$ et $`\overline{Q}`$ sont les mรชmes jusquโร lโordre $`k1`$ et sont nuls ร partir de lโordre $`k+1`$. Il sโagit donc simplement de trouver $`_k`$ et $`\overline{Q}_k`$. La condition $`Q=\overline{Q}`$ sโรฉcrit, en nรฉgligeant les signes provenant de la supersymรฉtrie :
$$_k(Q_1(x_1\mathrm{}x_k))+Q_k(x_1\mathrm{}x_k)=Q_1_k(x_1\mathrm{}x_k)+\overline{Q}_k(x_1\mathrm{}x_k)$$
$`()`$
oรน lโon a dรฉsignรฉ par la mรชme lettre $`Q_1`$ la dรฉrivation de lโalgรจbre $`S(\text{g}[1])`$ valant $`Q_1`$ sur $`\text{g}[1]`$. 1). Si tous les $`x_j,j=1\mathrm{}k`$ sont dans le noyau $`Z`$ de $`Q_1`$, lโรฉquation $`()`$ se rรฉduit ร :
$$Q_k(x_1\mathrm{}x_k)=Q_1_k(x_1\mathrm{}x_k)+\overline{Q}_k(x_1\mathrm{}x_k).$$
$`()_1`$
On choisit donc $`\overline{Q}_k(x_1\mathrm{}x_k)`$ comme รฉtant la projection de $`Q_k(x_1\mathrm{}x_k)`$ sur le supplรฉmentaire $`W\text{g}^{}`$ de $`B`$ dans g. Ceci permet de dรฉfinir $`_k(x_1\mathrm{}x_k)`$ ร un รฉlรฉment $`z`$ de $`Z`$ prรจs. On utilise alors lโรฉquation maรฎtresse $`[Q,Q]=0`$, qui permet de montrer, par rรฉcurrence sur $`k`$, que $`Q_k(x_1\mathrm{}x_k)`$ appartient ร $`Z`$. On en dรฉduit que $`\overline{Q}_k(x_1\mathrm{}x_k)`$ appartient bien ร $`\text{g}^{}`$. De plus si $`x_1=Q_1y_1B`$, lโรฉquation $`[Q,Q]=0`$ sโรฉcrit (toujours en nรฉgligeant les problรจmes de signes) :
$$Q_k(Q_1y_1.x_2\mathrm{}x_k)+\text{ termes intermรฉdiaires }+Q_1Q_k(y_1.x_2\mathrm{}x_k)=0.$$
Les termes intermรฉdiaires sont une somme de termes du type :
$$Q_j(\mathrm{}Q_l(\mathrm{})\mathrm{}),j,l<k.$$
Lโรฉlรฉment $`Q_1y_1`$ se trouve dans une parenthรจse intรฉrieure ou dans la parenthรจse extรฉrieure. Dans les deux cas lโhypothรจse de dรฉpart sur $`Q`$ entraรฎne lโannulation de ce terme. On a donc :
$$Q_k(Q_1y_1.x_2\mathrm{}x_k)=Q_1Q_k(y_1.x_2\mathrm{}x_k),$$
ce qui montre que $`\overline{Q}_k(Q_1y_1.x_2\mathrm{}x_k)=0`$. 2). Soit $`sS^k(\text{g}[1])`$, avec $`k2`$. On dit que $`x`$ est de type $`j,0jk`$, si $`x`$ sโรฉcrit $`x_1\mathrm{}x_k`$ avec $`x_1,\mathrm{},x_jW`$ et $`x_{j+1},\mathrm{},x_kZ`$. Nous allons dรฉterminer $`_k(x)`$ par rรฉcurrence (finie) sur le type de $`x`$, le type $`0`$ ayant รฉtรฉ traitรฉ au 1). On remarque que $`\overline{Q}_k(x)=0`$ si le type de $`x`$ est non nul. Lโรฉquation $`[Q,Q]=0`$ sโรฉcrit :
$$Q_k(Q_1(x_1\mathrm{}x_k))+Q_1Q_k(x_1\mathrm{}x_k)=0,$$
les termes intermรฉdiaires sโannulant pour la mรชme raison que dans le 1). On a donc :
$$Q_1Q_k(x)+Q_kQ_1(x)=0$$
$`(M)`$
pour tout $`xS^k(\text{g}[1])`$. Soit $`r1`$. Supposons que $`_k(x)`$ soit dรฉterminรฉ pour tout $`x`$ de type $`jr2`$, et dรฉterminรฉ ร un $`zZ`$ prรจs pour tout $`x`$ de type $`r1`$. Soit alors $`x`$ de type $`r`$. On veut dรฉterminer $`_k(x)`$ ร un รฉlรฉment $`z^{}Z`$ prรจs et prรฉciser $`_k(y)`$ pour tous les $`y`$ de type $`r1`$. Lโรฉquation $`()`$ appliquรฉe ร $`Q_1(x)`$ sโรฉcrit :
$$_kQ_1^2(x)+Q_kQ_1(x)=Q_1_kQ_1(x)$$
le terme $`\overline{Q}_kQ_1(x)`$ รฉtant nul. En reportant (M) dans cette รฉquation on a donc :
$$Q_1_kQ_1(x)+Q_1Q_k(x)=0,$$
dโoรน :
$$_kQ_1(x)+Q_k(x)Z.$$
Comme $`_kQ_1(x)`$ est dรฉterminรฉ ร un รฉlรฉment arbitraire de $`Z`$ prรจs, on peut sโarranger pour que :
$$_kQ_1(x)+Q_k(x)=b(x)$$
$`b(x)`$ appartient ร $`B`$. Lโรฉquation $`()`$ appliquรฉe ร $`x`$ sโรฉcrivant :
$$_kQ_1(x)+Q_k(x)=Q_1_k(x)$$
le choix dโun $`b(x)`$ nous permet de choisir $`_k(x)`$ ร un รฉlรฉment $`z^{}Z`$ prรจs. Le $`b(x)`$ doit obรฉir ร la contrainte suivante : si $`Q_1(x)=0`$, alors $`b(x)=Q_k(x)`$. Supposons que $`x=Q_1y`$$`y`$ est de type $`r+1`$. Alors, compte tenu de (M) la contrainte sur $`b`$ sโรฉcrit :
$$b(x)=Q_1Q_k(y).$$
Ayant choisi un $`b(x)`$ pour tout $`x`$ de type $`r`$ satisfaisant ร la contrainte ci-dessus, on peut alors choisir $`_k(x)`$ ร un รฉlรฉment $`z^{}Z`$ prรจs. Il reste donc simplement ร dรฉmontrer le lemme ci-dessous :
Lemme V.3.
Soit $`x`$ de type $`r1`$. Alors si $`Q_1x=0`$ il existe un $`y`$ de type $`r+1`$ tel que $`x=Q_1y`$.
Dรฉmonstration. On considรจre lโapplication $`\delta :\text{g}\text{g}`$ de degrรฉ $`1`$ dรฉfinie par $`\delta (x)=0`$ pour $`x\text{g}^{}W`$, et $`\delta (Q_1x)=x`$ pour tout $`x`$ dans g. On a alors :
$$Q_1\delta +\delta Q_1=\text{Id}p,$$
$`p`$ est la projection sur $`\text{g}^{}`$ parallรจlement ร $`\text{g}^{\prime \prime }`$ (autrement dit $`\delta `$ est une homotopie entre les deux endomorphismes de complexes $`\text{Id}`$ et $`p`$). Le lemme V.3 est un corollaire du rรฉsultat suivant, dรป ร Quillen \[Q appendix B\] :
Proposition V.4.
1). La dรฉrivation $`Q_1`$ de lโalgรจbre symรฉtrique $`S(\text{g})`$ vรฉrifie :
$$Q_1^2=0.$$
2). La cohomologie du complexe $`(S(\text{g}),Q_1)`$ est isomorphe ร $`S(\text{g}^{})`$, et un supplรฉmentaire de lโimage de $`Q_1`$ dans le noyau de $`Q_1`$ est donnรฉ par $`S(\text{g}^{})1`$ moyennant lโidentification : $`S(\text{g})=S(\text{g}^{})S(\text{g}^{\prime \prime })`$.
Dรฉmonstration.
1). Comme $`Q_1`$ est impaire, $`Q_1^2=\frac{1}{2}[Q_1,Q_1]`$ est encore une dรฉrivation de $`S(\text{g})`$. Comme $`Q_1^2{}_{|}{}^{}\text{g}=0`$ cette dรฉrivation est nulle. 2). On a : $`Q_1(v^{}v^{\prime \prime })=v^{}Q_1(v^{\prime \prime })`$ pour $`v^{}S(\text{g}^{})`$ et $`v^{\prime \prime }S(\text{g}^{\prime \prime })`$. On est ramenรฉ au cas oรน la cohomologie de g est triviale. On prolonge alors lโhomotopie $`\delta `$ ci-dessus en une dรฉrivation de $`S(\text{g})`$. On pose alors :
$$E=[Q_1,\delta ]=Q_1\delta +\delta Q_1.$$
$`E`$ est une dรฉrivation telle que $`E_|\text{g}=\text{Id}`$. On en dรฉduit :
$$E(x)=kx$$
pour tout $`xS^k(\text{g})`$. Si maintenant $`x`$ appartient ร $`S^k(\text{g})`$ et $`Q_1x=0`$, alors $`Ex=Q_1\delta x=kx`$. Si $`k1`$ on a donc :
$$x=Q_1(\frac{1}{k}\delta x).$$
La cohomologie de $`S(\text{g})`$ est donc rรฉduite au corps de base, qui est $`S(\{0\})`$.
$``$
Fin de la dรฉmonstration du lemme V.3 : le complexe $`V=S(\text{g})`$ admet ร son tour une dรฉcomposition :
$$V=V^{}V^{\prime \prime }$$
avec $`V^{}=S(\text{g}^{})1`$. Lโimage de $`Q_1`$ dans $`S(\text{g})`$ est lโidรฉal engendrรฉ par $`B=Q_1(\text{g})`$. On peut donc choisir pour $`V^{\prime \prime }`$ lโidรฉal engendrรฉ par $`\text{g}^{\prime \prime }`$. Le lemme provient alors du fait que tout รฉlรฉment de type $`r1`$ appartient ร cet idรฉal, sur lequel la cohomologie est triviale.
$``$
V.2. Dรฉmonstration du thรฉorรจme V.1 On se donne deux $`L_{\mathrm{}}`$-algรจbres $`(\text{g}_1,Q_1)`$ et $`(\text{g}_2,Q_2)`$ et un quasi-isomorphisme $``$ de $`(\text{g}_1,Q_1)`$ vers $`(\text{g}_2,Q_2)`$. Appliquant la proposition $`V.2`$ ร ces deux $`L_{\mathrm{}}`$-algรจbres on a le diagramme suivant, dans lequel toutes les flรจches sont des quasi-isomorphismes :
$$๐(\text{g}_{}^{}{}_{1}{}^{})\underset{i}{}๐(\text{g}_{}^{}{}_{1}{}^{}\text{g}_{}^{\prime \prime }{}_{1}{}^{})\stackrel{~}{}๐(\text{g}_1)\underset{}{}๐(\text{g}_2)\stackrel{~}{}๐(\text{g}_{}^{}{}_{2}{}^{}\text{g}_{}^{\prime \prime }{}_{2}{}^{})\underset{p}{\text{ }}๐(\text{g}_{}^{}{}_{2}{}^{}).$$
On a ainsi construit un quasi-isomorphisme $`^{}`$ entre deux $`L_{\mathrm{}}`$-algรจbres minimales. Son premier coefficient $`_1^{}:\text{g}_1^{}\text{g}_2^{}`$ รฉtant inversible, $`^{}`$ lui-mรชme est inversible. Lโajout du quasi-isomorphisme $`_{}^{}{}_{}{}^{1}`$ dans le diagramme ci-dessus permet alors la construction dโun quasi-isomorphisme :
$$๐ข:๐(\text{g}_2)๐(\text{g}_1)$$
qui est un quasi-inverse pour $``$.
$``$
VI. La formalitรฉ de Kontsevich Un $`L_{\mathrm{}}`$ morphisme entre $`T_{poly}\left(\text{}^d\right)`$ et $`D_{poly}\left(\text{}^d\right)`$ qui soit aussi un quasi-isomorphisme cโest ร dire un isomorphisme en cohomologie est une formalitรฉ.
M. Kontsevich a proposรฉ dans \[K1\] une formalitรฉ $`๐ฐ`$ explicite. Prรฉcisรฉment, les applications $`๐ฐ_n`$ sont donnรฉs par :
$$๐ฐ_n=\underset{m0}{}\underset{\stackrel{}{\mathrm{\Gamma }}G_{n,m}}{}w_\stackrel{}{\mathrm{\Gamma }}_\stackrel{}{\mathrm{\Gamma }}$$
$`G_{n,m}`$ est lโensemble des graphes orientรฉs admissibles ร $`n`$ sommets aรฉriens $`p_1`$,โฆ,$`p_n`$ et $`m`$ sommets terrestres $`q_1`$,โฆ,$`q_m`$ : de chaque sommet aรฉrien est issu $`k_1`$,โฆ, $`k_n`$ flรจches aboutissant soit ร un autre sommet aรฉrien soit ร un sommet terrestre. On ordonne les sommets aรฉriens et terrestres du graphe et on oriente le graphe en ordonnant les flรจches de faรงon compatible avec cet ordre, les flรจches issues du sommet $`p_j`$ ont les numรฉros $`k_1+\mathrm{}+k_{j1}+1`$,โฆ, $`k_1+\mathrm{}+k_j`$. On les note :
$$Star(p_j)=\{\stackrel{}{p_ja_1},\mathrm{},\stackrel{}{p_ja_{k_j}}\}\stackrel{}{v}_{k_1+\mathrm{}+k_{j1}+i}=\stackrel{}{p_ja_i}.$$
Si $`\stackrel{}{\mathrm{\Gamma }}`$ est un graphe orientรฉ, son poids $`w_\stackrel{}{\mathrm{\Gamma }}`$ est par dรฉfinition lโintรฉgrale sur lโespace de configuration $`C_{\{p_1,\mathrm{},p_n\},\{q_1,\mathrm{},q_m\}}^+`$ de la forme :
$$\omega _\stackrel{}{\mathrm{\Gamma }}=\frac{1}{(2\pi )^{{\scriptscriptstyle k_i}}k_1!\mathrm{}k_n!}d\mathrm{\Phi }_{\stackrel{}{v}_1}\mathrm{}d\mathrm{\Phi }_{\stackrel{}{v}_{k_1+\mathrm{}+k_n}}\text{}\mathrm{\Phi }_{\stackrel{}{p_ja}}=Arg\left(\frac{ap_j}{a\overline{p_j}}\right).$$
Enfin $`_\stackrel{}{\mathrm{\Gamma }}`$ est un opรฉrateur $`m`$-diffรฉrentiel, nul sur $`\alpha _1\text{.}\mathrm{}\text{.}\alpha _n`$ sauf si $`\alpha _1`$ est un $`k_1`$-tenseur, $`\alpha _2`$ un $`k_2`$-tenseur,โฆ, $`\alpha _n`$ un $`k_n`$-tenseur, auquel cas, on a :
$$_\stackrel{}{\mathrm{\Gamma }}(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n)(f_1,f_2,\mathrm{},f_m)=D_{p_1}\alpha _1^{i_1i_2\mathrm{}i_{k_1}}\mathrm{}D_{p_n}\alpha _n^{i_{k_1+\mathrm{}k_{n1}+1}\mathrm{}i_{k_1+\mathrm{}+k_n}}D_{q_1}f_1\mathrm{}D_{q_m}f_m$$
si $`D_a`$ est lโopรฉrateur :
$$D_a=\underset{l,\stackrel{}{v_l}=\stackrel{}{.a}}{}_{i_l}$$
et si la somme est รฉtendue ร tous les indices $`i_j`$ rรฉpรฉtรฉs. On notera aussi
$$๐ฐ_n=๐ฐ_{(k_1,k_2,\mathrm{},k_n)}=๐ฐ_{k_{\{1,\mathrm{},n\}}}.$$
Maintenant, si on change lโordre des flรจches issues dโun sommet $`p_j`$, le produit $`w_\stackrel{}{\mathrm{\Gamma }}_\stackrel{}{\mathrm{\Gamma }}`$ ne change pas. On prend la convention suivante : si $`\stackrel{}{\mathrm{\Gamma }}`$ est un graphe orientรฉ de faรงon non compatible, on pose :
$$B_\stackrel{}{\mathrm{\Gamma }}=\epsilon (\sigma )B_{\stackrel{}{\mathrm{\Gamma }}^\sigma }$$
$`\sigma `$ est nโimporte quelle permutation des flรจches de $`\stackrel{}{\mathrm{\Gamma }}`$ qui le transforme en un graphe $`\stackrel{}{\mathrm{\Gamma }}^\sigma `$ orientรฉ de faรงon compatible. Avec cette convention, on aura :
$$๐ฐ_n=\underset{m0}{}\underset{\stackrel{}{\mathrm{\Gamma }}G_{n,m}^{}}{}w_\stackrel{}{\mathrm{\Gamma }}^{}_\stackrel{}{\mathrm{\Gamma }}^{}$$
$`G_{n,m}^{}`$ est lโensemble de tous les graphes orientรฉs de faรงon compatible ou non et $`w_\stackrel{}{\mathrm{\Gamma }}^{}`$ est lโintรฉgrale de la forme :
$$\omega _\stackrel{}{\mathrm{\Gamma }}^{}=\frac{1}{(2\pi )^{{\scriptscriptstyle k_i}}(k_i)!}d\mathrm{\Phi }_{\stackrel{}{v}_1}\mathrm{}d\mathrm{\Phi }_{\stackrel{}{v}_{k_1+\mathrm{}+k_n}}\text{}\mathrm{\Phi }_{\stackrel{}{p_ja}}=Arg\left(\frac{ap_j}{a\overline{p_j}}\right).$$
Nous allons vรฉrifier dans la suite que nos choix de signes sont cohรฉrents.
Thรฉorรจme VI.1 (M.Kontsevich).
Lโapplication formelle $`๐ฐ`$ est une formalitรฉ. En particulier cโest un $`L_{\mathrm{}}`$-morphisme.
Dรฉmonstration. Puisque $`Q_1=0`$, lโรฉquation de formalitรฉ sโรฉcrit :
$$\begin{array}{cc}\hfill 0=& Q_1^{}\left(๐ฐ_{k_{\{1,\mathrm{},n\}}}(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n)\right)+\hfill \\ & +\frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (I,J)Q_2^{}\left(๐ฐ_{k_I}(\alpha _I)๐ฐ_{k_J}(\alpha _J)\right)\hfill \\ & \frac{1}{2}\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left(Q_2(\alpha _i\text{.}\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right).\hfill \end{array}$$
Remarquons maintenant que pour que $`w_\stackrel{}{\mathrm{\Gamma }}`$ ne soit pas nul, il faut que le degrรฉ de la forme $`\omega _\stackrel{}{\mathrm{\Gamma }}`$ soit รฉgal ร la dimension de lโespace de configuration $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ sur lequel on intรจgre, cโest ร dire :
$$k_i=2n+m2.$$
Dans ce cas,
$$(1)^m=(1)^{\left|๐ฐ_{k_{\{1,\mathrm{},n\}}}(\mathrm{})\right|+1}=(1)^{{\scriptscriptstyle k_i}}=(1)^{|k_{\{\}1,\mathrm{}n}|}.$$
Donc notre รฉquation devient :
$$\begin{array}{cc}\hfill (1)& 0=๐ฐ_{k_{\{1,\mathrm{},n\}}}(\alpha _{\{1,\mathrm{},n\}})\mu (1)^{{\scriptscriptstyle k_i}1}\mu ๐ฐ_{k_{\{1,\mathrm{},n\}}}(\alpha _{\{1,\mathrm{},n\}})+\hfill \\ \hfill (2)& +\frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (I,J)(1)^{(|k_I|1)|k_J|}๐ฐ_{k_I}(\alpha _I)๐ฐ_{k_J}(\alpha _J)+\hfill \\ \hfill (3)& +\frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (I,J)(1)^{|k_I|}๐ฐ_{k_J}(\alpha _J)๐ฐ_{k_I}(\alpha _I)\hfill \\ \hfill (4)& \frac{1}{2}\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left((\alpha _i\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \\ \hfill (5)& \frac{1}{2}\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)๐ฐ_{((k_i+k_j1),\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left((1)^{k_ik_j}(\alpha _j\alpha _i)\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right).\hfill \end{array}$$
Montrons que $`(2)=(3)`$. En fait :
$$\epsilon _\alpha (I,J)=\epsilon _\alpha (J,I)(1)^{|k_I||k_J|}$$
car le nombre de $`i`$ de $`I`$ tel que $`k_i2`$ soit impair est congru modulo ร 2 ร $`|k_I|=k_i`$. Donc :
$$(3)=\frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (J,I)(1)^{|k_I||k_J|+|k_I|}๐ฐ_{k_J}(\alpha _J)๐ฐ_{k_I}(\alpha _I)$$
en changeant les rรดles de $`I`$ et $`J`$ :
$$(3)=\frac{1}{2}\underset{\genfrac{}{}{0pt}{}{IJ=\{1,\mathrm{},n\}}{I,J\mathrm{}}}{}\epsilon _\alpha (I,J)(1)^{|k_J|(|k_I|1)}๐ฐ_{k_I}(\alpha _I)๐ฐ_{k_J}(\alpha _J)=(2).$$
De mรชme $`(5)=(4)`$ :
$$\begin{array}{cc}& (4)=\frac{1}{2}\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left((\alpha _i\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \\ & =\frac{1}{2}\underset{ij}{}\epsilon _\alpha (j,i,1,\mathrm{},\widehat{i,j},\mathrm{},n)(1)^{k_ik_j}๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left((\alpha _i\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \\ & =\frac{1}{2}\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_j},\mathrm{},\widehat{k_i},\mathrm{},k_n)}\left((1)^{k_ik_j}(\alpha _j\alpha _i)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \\ & =(5).\hfill \end{array}$$
Posons enfin $`\mu =๐ฐ_{\mathrm{}}`$. Alors $`(1)`$ sโรฉcrit :
$$(1)=(1)^{(|k_{\{1,\mathrm{},n\}}|1).0}๐ฐ_{k_{\{1,\mathrm{},n\}}}\left(\alpha _{\{1,\mathrm{},n\}}\right)๐ฐ_{\mathrm{}}+(1)^{(01)|k_{\{1,\mathrm{},n\}}|}๐ฐ_{\mathrm{}}๐ฐ_{k_{\{1,\mathrm{},n\}}}\left(\alpha _{k_{\{1,\mathrm{},n\}}}\right).$$
Comme $`\epsilon _\alpha (\{1,\mathrm{},n\},\mathrm{})=\epsilon _\alpha (\mathrm{},\{1,\mathrm{},n\})=1`$, lโรฉquation de formalitรฉ devient :
$$\begin{array}{cc}\hfill \underset{IJ=\{1,\mathrm{},n\}}{}& \epsilon _\alpha (I,J)(1)^{(|k_I|1)|k_J|}๐ฐ_{k_I}(\alpha _I)๐ฐ_{k_J}(\alpha _J)\underset{ij}{}\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)\hfill \\ & ๐ฐ_{((k_i+k_j1),k_1,\mathrm{},\widehat{k_i},\mathrm{},\widehat{k_j},\mathrm{},k_n)}\left((\alpha _i\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right)=0.\hfill \end{array}$$
Si on remplace les $`๐ฐ_I(\alpha _I)`$ par les $`_\stackrel{}{\mathrm{\Gamma }}w_\stackrel{}{\mathrm{\Gamma }}^{}_\stackrel{}{\mathrm{\Gamma }}^{}(\alpha _I)`$ et quโon dรฉveloppe tout, on obtient une somme dโopรฉrateurs multi-diffรฉrentiels de la forme :
$$\underset{\stackrel{}{\mathrm{\Gamma }^{}}}{}c_\stackrel{}{\mathrm{\Gamma }^{}}_\stackrel{}{\mathrm{\Gamma }^{}}^{}(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n)$$
$`\stackrel{}{\mathrm{\Gamma }^{}}`$ est un graphe ร $`n`$ sommets aรฉriens, $`m`$ sommets terrestres ayant $`2n+m3`$ flรจches. Si on se donne $`\stackrel{}{\mathrm{\Gamma }^{}}`$ orientรฉ et une face $`F`$ de codimension 1 de $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$, on associe ร ce couple $`(\mathrm{\Gamma }^{},F)`$ au plus un terme de lโรฉquation de formalitรฉ. Plus prรฉcisรฉment :
Cas 1 : si
$$\begin{array}{cc}\hfill F& =_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_{l+1},\mathrm{},q_{l+m_1}\}}C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+\hfill \\ & =C_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_{l+1},\mathrm{},q_{l+m_1}\}}^+\times C_{\{p_1,\mathrm{},p_n\}\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_1,\mathrm{},q_l,q,q_{l+m_1+1},\mathrm{},q_m\}}^+,\hfill \end{array}$$
que lโon notera :
$$_{S,S^{}}C_{A,B}^+=C_{S,S^{}}^+\times C_{AS,BS^{}\{q\}}^+,$$
on associe au couple $`(\stackrel{}{\mathrm{\Gamma }}^{},F)`$ lโunique terme :
$$_{\stackrel{}{\mathrm{\Gamma }}^{},F}^{}(\alpha _1,..,\alpha _n)(f_1,\mathrm{},f_m)=_{\stackrel{}{\mathrm{\Gamma }}_2}^{}\left(\alpha _{j_1}\text{.}\mathrm{}\text{.}\alpha _{j_{n_2}}\right)(f_1,\mathrm{},f_l,_{\stackrel{}{\mathrm{\Gamma }}_1}^{}\left(\alpha _{i_1}\text{.}\mathrm{}\text{.}\alpha _{i_{n_1}}\right)(f_{l+1},\mathrm{},f_{l+m_1}),f_{l+m_1+1},\mathrm{}f_m)$$
$`\stackrel{}{\mathrm{\Gamma }}_1`$ est la restriction ร $`\{p_{i_1},\mathrm{},p_{i_{n_1}}\}\{q_{l+1},\mathrm{},q_{l+m_1}\}`$ (avec son ordre), $`\stackrel{}{\mathrm{\Gamma }}_2`$ est le graphe obtenu en collapsant les points $`p_{i_1}`$,โฆ, $`p_{i_{n_1}}`$ et $`q_{l+1}`$,โฆ, $`q_{l+m_1}`$ en $`q`$, on a posรฉ $`\{1,\mathrm{},n\}\{i_1,\mathrm{},i_{n_1}\}=\{j_1<j_2<\mathrm{}<j_{n_2}\}`$. On note $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}`$ le coefficient de cet opรฉrateur.
Remarquons que lโapplication $`(\stackrel{}{\mathrm{\Gamma }}^{},F)(\stackrel{}{\mathrm{\Gamma }}_1,\stackrel{}{\mathrm{\Gamma }}_2)`$ est dans ce cas surjective mais pas injective. Si on se donne le couple $`(\stackrel{}{\mathrm{\Gamma }}_1,\stackrel{}{\mathrm{\Gamma }}_2)`$, la face $`F`$ est bien dรฉterminรฉe mais $`\stackrel{}{\mathrm{\Gamma }^{}}`$ nโest pas unique : il y a dโabord la rรฉpartition des flรจches allant dโun sommet de $`\stackrel{}{\mathrm{\Gamma }}_2`$ vers un sommet de $`\stackrel{}{\mathrm{\Gamma }}_1`$ (application de la rรจgle de Leibniz) chaque rรฉpartition correspond ร un graphe $`\mathrm{\Gamma }_2`$ diffรฉrent. Si cette rรฉpartition est donnรฉe, il faut encore fixer lโordre des flรจches de $`\mathrm{\Gamma }^{}`$. Le nombre de choix est bien sรปr le quotient du nombre dโorientations possibles pour $`\mathrm{\Gamma }^{}`$ par celui des orientations possibles de $`\mathrm{\Gamma }_1`$ et $`\mathrm{\Gamma }_2`$ :
$$\text{Nombre dโorientations de }\mathrm{\Gamma }^{}=\frac{\left(k_{i_1}+\mathrm{}+k_{i_{n_1}}\right)!\left(k_{j_1}+\mathrm{}+k_{j_{n_2}}\right)!}{\left(k_i\right)!}=\frac{|k_I|!|k_J|!}{\left|k_{\{1,\mathrm{},n\}}\right|!}.$$
Cas 2 : si
$$F=_{\{p_i,p_j\}}C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+=C_{\{p_i,p_j\}}\times C_{\{p,p_1,\mathrm{},\widehat{p_i},\mathrm{},\widehat{p_j},\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+,$$
que lโon notera :
$$_SC_{A,B}^+=C_S\times C_{AS\{p\},B}^+,$$
A $`(\stackrel{}{\mathrm{\Gamma }}^{},F)`$, si la flรจche $`\stackrel{}{p_ip_j}`$ est une des flรจches de $`\mathrm{\Gamma }^{}`$, on associe lโunique terme :
$$_{\stackrel{}{\mathrm{\Gamma }}^{},F}^{}(\alpha _1,..,\alpha _n)(f_1,\mathrm{},f_m)=_{\stackrel{}{\mathrm{\Gamma }}_2}\left((\alpha _i\alpha _j)\text{.}\alpha _1\text{.}\mathrm{}\text{.}\widehat{\alpha _i}\text{.}\mathrm{}\text{.}\widehat{\alpha _j}\text{.}\mathrm{}\text{.}\alpha _n\right)$$
$`\stackrel{}{\mathrm{\Gamma }}_2`$ est le graphe obtenu en collapsant les sommets $`p_i`$ et $`p_j`$ du graphe $`\mathrm{\Gamma }^{}`$ sur le point $`p`$ et en รฉliminant la flรจche $`\stackrel{}{p_ip_j}`$. Si cette flรจche nโexiste pas dans $`\mathrm{\Gamma }^{}`$, on associe lโopรฉrateur nul ร $`(\stackrel{}{\mathrm{\Gamma }}^{},F)`$. On note $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}`$ le coefficient de cet opรฉrateur.
Dans ce cas, on considรฉrera lโapplication $`(\stackrel{}{\mathrm{\Gamma }}^{},F)(\stackrel{}{\mathrm{\Gamma }}_1,\stackrel{}{\mathrm{\Gamma }}_2)`$$`\stackrel{}{\mathrm{\Gamma }}_1`$ est le graphe tracรฉ dans $`C_{\{p_i,p_j\}}`$ ร une seule flรจche : la flรจche $`\stackrel{}{p_ip_j}`$. A part le cas 0, lโimage rรฉciproque dโun couple $`(\stackrel{}{\mathrm{\Gamma }}_1,\stackrel{}{\mathrm{\Gamma }}_2)`$ contient exactement :
$$\text{Nombre dโorientations de }\mathrm{\Gamma }^{}=\frac{\left((k_i+k_j1)+k_1+\mathrm{}+\widehat{k_i}+\mathrm{}+\widehat{k_j}+\mathrm{}+k_n\right)!}{\left(k_i\right)!}=\frac{\left(|k_{\{1,\mathrm{},n\}}|1\right)!}{\left|k_{\{1,\mathrm{},n\}}\right|!}.$$
Cas 3 : si
$$F=_SC_{A,B}^+=C_S\times C_{AS\{p\},B}^+$$
avec $`|S|3`$, dans ce cas aucun terme de lโรฉquation de formalitรฉ nโest associรฉ ร $`(\stackrel{}{\mathrm{\Gamma }}^{},F)`$. On pose donc $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}=0`$.
Pour chaque $`\stackrel{}{\mathrm{\Gamma }}^{}`$, on dรฉfinit sur $`C_{\{p_1,\mathrm{},p_n\};\{q_1,\mathrm{},q_m\}}^+`$ la forme :
$$\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}=\frac{1}{(2\pi )^{\left|k_{\{1,\mathrm{},n\}}\right|}\left|k_{\{1,\mathrm{},n\}}\right|!}d\mathrm{\Phi }_{\stackrel{}{v}_1}\mathrm{}d\mathrm{\Phi }_{\stackrel{}{v}_{k_1+\mathrm{}+k_n}}.$$
Montrons quโavec toutes ces notations, lโรฉquation de formalitรฉ sโรฉcrit :
$$\begin{array}{cc}\hfill 0& =\underset{\stackrel{}{\mathrm{\Gamma }}^{}G_{n,m}^{}}{}\left[\underset{FC_{A,B}^+}{}_\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}\right]_\stackrel{}{\mathrm{\Gamma }}^{}^{}\left(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n\right)\hfill \\ & =\underset{\stackrel{}{\mathrm{\Gamma }}^{}G_{n,m}^{}}{}\left[_{C_{A,B}^+}๐\omega _\stackrel{}{\mathrm{\Gamma }}^{}\right]_\stackrel{}{\mathrm{\Gamma }}^{}^{}\left(\alpha _1\text{.}\mathrm{}\text{.}\alpha _n\right).\hfill \end{array}$$
Le rรฉsultat est donc une simple consรฉquence du thรฉorรจme de Stokes sur la variรฉtรฉ ร coins $`C_{A,B}^+`$ et pour les formes fermรฉes $`\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}`$.
Comparons donc terme par terme chaque coefficient $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}`$ et lโintรฉgrale sur la face orientรฉe $`\stackrel{}{F}`$ de la forme $`\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}`$.
Cas 1 : Avec nos notations, le coefficient $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}`$ est :
$$\begin{array}{cc}\hfill c_{\stackrel{}{\mathrm{\Gamma }}^{},F}& =\epsilon _\alpha (J,I)(1)^{(|k_J|1)|k_I|}\frac{|k_I|!|k_J|!}{\left|k_{\{1,\mathrm{},n\}}\right|!}(1)^{l(m_11)}\hfill \\ & _{C_{S,S^{}}^+}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}_{C_{AS,BS^{}\{q\}}^+}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{}.\hfill \end{array}$$
(Le signe $`(1)^{l(m_11)}`$ provient du dรฉveloppement de lโopรฉration $``$). On rappelle que $`S^{}`$ est le segment $`q_{l+1},\mathrm{},q_{l+m_1}`$ et que $`q`$ remplace $`S^{}`$ dans $`BS^{}\{q\}`$. Dโautre part, on a pour la forme
$$\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}=\epsilon _\alpha (I,J)\frac{|k_I|!|k_J|!}{\left|k_{\{1,\mathrm{},n\}}\right|!}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{}$$
et la face $`\stackrel{}{F}`$ (en tenant compte de son orientation) :
$$\begin{array}{cc}\hfill _\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}& =\epsilon _\alpha (I,J)(1)^{lm_1+l+m_1}\frac{|k_I|!|k_J|!}{\left|k_{\{1,\mathrm{},n\}}\right|!}_{C_{\{p_{i_1},\mathrm{},p_{i_{n_1}}\};\{q_{l+1},\mathrm{},q_{l+m_1}\}}^+}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}\hfill \\ & _{C_{AS,BS^{}\{q\}}^+}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{}.\hfill \end{array}$$
Rappelons que $`|k_I|=2n_1+m_12`$, $`|k_J|=2(nn_1)+(mm_1+1)2`$, le signe devant lโintรฉgrale est donc :
$$\begin{array}{cc}\hfill \epsilon _\alpha (I,J)(1)^{lm_1+l+m_1}& =\epsilon _\alpha (J,I)(1)^{|k_I||k_J|}(1)^{lm_1+m_1+l}\hfill \\ & =\epsilon _\alpha (J,I)(1)^{|k_I||k_J|}(1)^{lm_1+l}(1)^{|k_I|}.\hfill \end{array}$$
Donc :
$$c_{\stackrel{}{\mathrm{\Gamma }}^{},F}=_\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}.$$
Cas 2 : Avec nos notations, le coefficient $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}`$ est nul si $`\stackrel{}{\mathrm{\Gamma }}^{}`$ ne contient pas la flรจche $`\stackrel{}{p_ip_j}`$ et sinon :
$$\begin{array}{cc}\hfill c_{\stackrel{}{\mathrm{\Gamma }}^{},F}& =\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)\frac{(|k_{\{1,\mathrm{},n\}}|1)!}{\left|k_{\{1,\mathrm{},n\}}\right|!}\hfill \\ & _{C_S}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}_{C^+AS\{p\},B}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{}.\hfill \end{array}$$
Dโautre part, on a pour la forme
$$\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}=\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)\frac{(|k_{\{1,\mathrm{},n\}}|1)!}{\left|k_{\{1,\mathrm{},n\}}\right|!}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{}$$
et la face $`\stackrel{}{F}`$ (en tenant compte de son orientation) :
$$_\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}=\epsilon _\alpha (i,j,1,\mathrm{},\widehat{i,j},\mathrm{},n)\frac{(|k_{\{1,\mathrm{},n\}}|1)!}{\left|k_{\{1,\mathrm{},n\}}\right|!}_{C_S}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}_{C_{AS\{p\},B}^+}\omega _{\stackrel{}{\mathrm{\Gamma }}_2}^{},$$
avec $`S=\{p_i,p_j\}`$. On a donc pour tout $`\stackrel{}{\mathrm{\Gamma }}^{}`$ et $`F`$ :
$$c_{\stackrel{}{\mathrm{\Gamma }}^{},F}=_\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}.$$
Cas 3 : Il nโy a pas de termes dans notre รฉquation de formalitรฉ dans ce cas, ou $`c_{\stackrel{}{\mathrm{\Gamma }}^{},F}=0`$. Mais dans ce cas, on a le lemme suivant de Kontsevich \[K1 ยง 6.6.1\], \[Kh\] :
$$_{C_S}\omega _{\stackrel{}{\mathrm{\Gamma }}_1}^{}=0$$
si $`|S|3`$. On a donc de nouveau :
$$c_{\stackrel{}{\mathrm{\Gamma }}^{},F}=_\stackrel{}{F}\omega _\stackrel{}{\mathrm{\Gamma }}^{}^{}.$$
Et ceci finit la preuve de la validitรฉ de lโรฉquation de formalitรฉ.
$``$
* *
*
Appendice. Formalitรฉ et quantification par dรฉformation Nous expliquons dans ce paragraphe pourquoi la formalitรฉ de Kontsevich permet dโobtenir un รฉtoile-produit ร partir dโun $`2`$-tenseur de Poisson. On considรจre une (limite projective dโ) algรจbre(s) nilpotente(s) de dimension finie m. Par exemple :
$$\text{m}=\mathrm{}\mathrm{}[[\mathrm{}]]=\underset{}{lim}\mathrm{}\mathrm{}[[\mathrm{}]]/\mathrm{}^k\mathrm{}[[\mathrm{}]].$$
A.1.1 Construction dโรฉtoile-produits On se donne une $`L_{\mathrm{}}`$-algรจbre $`(\text{g},Q)`$ sur un corps $`k`$ de caractรฉristique zรฉro, que lโon voit comme une $`Q`$-variรฉtรฉ formelle graduรฉe pointรฉe. Un m-point de la variรฉtรฉ formelle g est par dรฉfinition un morphisme de cogรจbres :
$$p:\text{m}^{}๐(\text{g}).$$
Le produit tensoriel (complรฉtรฉ dans le cas dโune limite projective) $`๐(\text{g})\widehat{}\text{m}`$, muni de la comultiplication de $`๐(\text{g})`$ รฉtendue par m-linรฉaritรฉ, admet une structure de cogรจbre (sans co-unitรฉ) sur m. On peut alors voir un m-point comme un รฉlรฉment non nul de type groupe de cette cogรจbre, cโest-ร -dire un รฉlรฉment $`p๐(\text{g})\widehat{}\text{m}`$ vรฉrifiant : $`\mathrm{\Delta }p=pp`$.
Proposition A.1.
Les m-points de la variรฉtรฉ formelle g sont donnรฉs par :
$$p_v=e^v1=v+\frac{v^2}{2}+\mathrm{}$$
$`v`$ est un รฉlรฉment pair de $`\text{g}[1]\widehat{}\text{m}`$
Dรฉmonstration. La sรฉrie a bien un sens dans $`๐(\text{g})\widehat{}\text{m}`$. Si $`p`$ est un m-point, $`p`$ est forcรฉment pair, et on voit que la sรฉrie :
$$v=\text{Log}(1+p)=p\frac{p^2}{2}+\mathrm{}$$
a un sens dans $`๐(\text{g})\widehat{}\text{m}`$ et dรฉfinit un รฉlรฉment primitif (et pair), cโest-ร -dire que lโon a : $`\mathrm{\Delta }v=0`$. Pour dรฉmontrer ce point on rajoute formellement la co-unitรฉ en considรฉrant la cogรจbre :
$$\overline{๐}_\text{m}=(k.1\text{m})๐(\text{g})\widehat{}\text{m}.$$
Un รฉlรฉment de type groupe de cette cogรจbre sโรฉcrit toujours :
$$g=1+p$$
$`p`$ est de type groupe dans la cogรจbre sans co-unitรฉ $`๐(\text{g})\widehat{}\text{m}`$. Il sโagit alors de montrer que le logarithme $`v`$ dโun tel รฉlรฉment est primitif, cโest-ร -dire que lโon a dans $`\overline{๐}_\text{m}`$ :
$$\mathrm{\Delta }v=v1+1v$$
Pour cela on remarque que la cogรจbre $`\overline{๐}_\text{m}`$ est en fait une bigรจbre. Le calcul formel suivant a alors un sens :
$$\begin{array}{cc}\hfill \mathrm{\Delta }\text{Log}g& =\text{Log}(\mathrm{\Delta }g)\hfill \\ & =\text{Log}(gg)\hfill \\ & =\text{Log}\left((g1)(1g)\right)\hfill \\ & =\text{Log}(g1)+\text{Log}(1g)\hfill \\ & =\text{Log}g1+1\text{Log}g\hfill \end{array}$$
Donc, forcรฉment $`v`$ appartient ร $`\text{g}\widehat{}\text{m}`$, et il est clair que $`p=p_v`$.
$``$
Le champ de vecteurs $`Q`$ sโรฉtend de maniรจre naturelle ร $`๐(\text{g})\widehat{}\text{m}`$. Supposons que $`Q`$ sโannule au point $`p_v`$ :
$$Q(e^v1)=0$$
On traduit ceci par le fait que $`v`$ vรฉrifie lโรฉquation de Maurer-Cartan gรฉnรฉralisรฉe :
$$Q_1(v)+\frac{1}{2}Q_2(v.v)+\mathrm{}=0.$$
(MCG)
Si g est une algรจbre de Lie diffรฉrentielle graduรฉe, รงa se rรฉduit ร lโรฉquation de Maurer-Cartan :
$$dv\frac{1}{2}[v,v]=0.$$
(En effet $`v`$ est pair dans $`\text{g}[1]\text{m}`$, donc impair dans $`\text{g}\text{m}`$). Si maintenant $``$ est un $`L_{\mathrm{}}`$-morphisme entre $`(\text{g}_1,Q)`$ et $`(\text{g}_2,Q^{})`$, et si $`v\text{g}\text{m}`$ est tel que $`Q(p_v)=0`$, il est clair que :
$$Q^{}\left((p_v)\right)=\left(Q(p_v)\right)=0.$$
Or $`(p_v)=e^w1`$ avec $`w\text{g}_2\widehat{}\text{m}`$ dโaprรจs la proposition A.1, puisque $`(p_v)`$ est de type groupe. Il est clair que $`w`$ est la projection canonique de $`(p_v)`$ sur $`\text{g}_2\widehat{}m`$, soit :
$$\begin{array}{cc}\hfill w& =\underset{n1}{}\frac{1}{n!}_n(v^n).\hfill \end{array}$$
En rรฉsumรฉ, si $`v\text{g}_1\widehat{}\text{m}`$ vรฉrifie (MCG), alors lโรฉlรฉment $`w\text{g}_2\widehat{}\text{m}`$ donnรฉ par lโรฉgalitรฉ ci-dessus vรฉrifie (MCG). Dans le cas oรน les deux $`L_{\mathrm{}}`$-algรจbres sont les algรจbres de Lie diffรฉrentielles graduรฉes des multichamps de vecteurs et des opรฉrateurs polydiffรฉrentiels, tout $`2`$-tenseur de Poisson formel :
$$v=\mathrm{}\gamma _1+\mathrm{}^2\gamma _2+\mathrm{}$$
donne naissance grรขce ร ce processus ร un opรฉrateur bidiffรฉrentiel formel $`w`$ tel que $`\mu +w`$ soit un รฉtoile-produit, $`\mu `$ dรฉsignant la multiplication usuelle de deux fonctions. A.2. Equivalence des foncteurs de dรฉformation On suppose toujours le corps de base $`k`$ de caractรฉristique zรฉro. Soit g une algรจbre de Lie diffรฉrentielle graduรฉe. Rappelons \[K1 ยง 3.2\] que le foncteur de dรฉformation $`\text{Def}_\text{g}`$ associe ร toute algรจbre commutative nilpotente de dimension finie m lโensemble des classes de solutions de degrรฉ $`1`$ de lโรฉquation de Maurer-Cartan dans $`\text{g}\text{m}`$ modulo lโaction du groupe de jauge, cโest-ร -dire le groupe nilpotent $`G_\text{m}=\mathrm{exp}(\text{g}^0\text{m})`$, dont lโaction (par des transformations affines de lโespace $`\text{g}^1\text{m}`$) est donnรฉe infinitรฉsimalement par :
$$\alpha .\gamma =d\alpha +[\alpha ,\gamma ]$$
pour tout $`\alpha \text{g}^0\text{m}`$ et pour tout $`\gamma \text{g}^1\text{m}`$. Ce foncteur sโรฉtend naturellement aux limites projectives dโalgรจbres commutatives nilpotentes de dimension finie : $`\text{Def}_\text{g}(\text{m})`$ est dans ce cas dรฉfini comme lโensemble des classes de solutions de degrรฉ $`1`$ de lโรฉquation de Maurer-Cartan dans le produit tensoriel complรฉtรฉ $`\text{g}\widehat{}\text{m}`$ modulo lโaction du groupe pro-nilpotent $`G_\text{m}=\mathrm{exp}(\text{g}^0\widehat{}\text{m})`$. Lโรฉquivalence de jauge peut aussi se dรฉfinir pour une $`L_{\mathrm{}}`$-algรจbre quelconque $`(\text{g},Q)`$ : deux solutions de lโรฉquation de Maurer-Cartan gรฉnรฉralisรฉe $`\gamma _0`$ et $`\gamma _1`$ dans $`\text{g}^1\text{m}`$ sont รฉquivalentes sโil existe une famille polynomiale $`\xi (t)_{tk}`$ de champs de vecteurs de degrรฉ $`1`$ et une famille polynomiale $`\gamma (t)_{tk}`$ de solutions de lโรฉquation de Maurer-Cartan gรฉnรฉralisรฉe dans $`\text{g}^1\text{m}`$ telles que :
$$\begin{array}{cc}\hfill \frac{d\gamma (t)}{dt}& =[Q,\xi (t)]\left(\gamma (t)\right)\hfill \\ \hfill \gamma (0)& =\gamma _0,\gamma (1)=\gamma _1.\hfill \end{array}$$
$`()`$
On vรฉrifie facilement que cette relation est une relation dโรฉquivalence, ce qui permet de dรฉfinir le foncteur de dรฉformation $`\text{Def}_\text{g}`$ comme la correspondance qui ร toute algรจbre commutative nilpotente de dimension finie m associe lโensmble $`\text{Def}_\text{g}(\text{m})`$ des classes de solutions de degrรฉ $`1`$ de lโรฉquation de Maurer-Cartan gรฉnรฉralisรฉe dans $`\text{g}\text{m}`$ modulo lโรฉquivalence de jauge.
Proposition A.2.1 (cf. \[K1 ยง 4.5.2\]).
1). Dans le cas dโune algรจbre de Lie diffรฉrentielle graduรฉe les deux notions dโรฉquivalence de jauge (et donc de foncteur de dรฉformation) coรฏncident. 2). Soient $`\text{g}_1`$ et $`\text{g}_2`$ deux algรจbres de Lie diffรฉrentielles graduรฉes. Alors le foncteur $`\text{Def}_{\text{g}_1\text{g}_2}`$ est naturellement รฉquivalent au produit des foncteurs $`\text{Def}_{\text{g}_1}\times \text{Def}_{\text{g}_2}`$. 3). Le foncteur de dรฉformation est trivial pour une $`L_{\mathrm{}}`$-algรจbre linรฉaire contractile.
Dรฉmonstration. Les points 2) et 3) sont faciles ร รฉtablir. Pour รฉtablir le premier point on remarque que lโaction de $`\alpha \text{g}^0\text{m}`$ sur $`\text{g}^1\text{m}`$ est donnรฉe par le champ de vecteurs :
$$D_\alpha =[Q,R_\alpha ],$$
$`R_\alpha `$ est le champ de vecteurs constant รฉgal ร $`\alpha `$. Ce champ de vecteurs est bien de degrรฉ $`1`$. Lโรฉquivalence de jauge au sens des algรจbres de Lie diffรฉrentielles graduรฉes $`\gamma _1=(\mathrm{exp}\alpha ).\gamma _0`$ entraรฎne donc lโรฉquivalence de jauge au sens des $`L_{\mathrm{}}`$-algรจbres, avec $`\xi (t)=R_\alpha `$ pour tout $`t`$ et $`\gamma (t)=(\mathrm{exp}t\alpha ).\gamma _0`$. Supposons maintenant que $`\gamma _0`$ et $`\gamma _1`$ sont รฉquivalents au sens des $`L_{\mathrm{}}`$-algรจbres. Soient $`\xi (t)`$ et $`\gamma (t)`$ les familles polynomiales de champs de vecteurs de degrรฉ $`1`$ et de solutions de lโรฉquation de Maurer-Cartan gรฉnรฉralisรฉe respectivement, telles que lโรฉquation (\**) soit vรฉrifiรฉe. Le champ de vecteurs de degrรฉ zรฉro $`[Q,\xi (t)]`$ sโรฉcrit explicitement en tout m-point $`\gamma \text{g}^1\text{m}`$ :
$$[Q,\xi (t)](\gamma )=d\xi (t)+[\xi (t),\gamma ].$$
On effectue ce calcul en appliquant la codรฉrivation $`[Q,\xi (t)]`$ ร lโรฉlรฉment de type groupe $`e^\gamma 1`$. Compte tenu de lโรฉquation (\**) on voit que le vecteur tangent $`\frac{d\gamma (t)}{dt}`$ au point $`\gamma (t)`$ est donnรฉ par un champ de vecteurs provenant de lโaction dโun รฉlรฉment de lโalgรจbre de Lie $`\text{g}^0\text{m}`$. Montrons par rรฉcurrence sur le degrรฉ $`d`$ de $`\gamma _t`$ (en tant que polynรดme) quโil existe un entier $`r`$ (dรฉpendant de $`d`$) tel que pour tout $`tk`$ il existe $`g_t=\mathrm{exp}(tD_1+\mathrm{}+t^rD_r)G_\text{m}`$ tel que $`\gamma _t=g_t.\gamma _0`$ : si $`d=0`$ on a $`\gamma _t=\gamma _0`$ et la constante $`g_t=\text{Id}`$ convient. Supposons donc que la propriรฉtรฉ soit vraie au rang $`d1`$. Supposons que $`\gamma _t`$ soit un polynรดme de degrรฉ $`d`$, que lโon peut รฉcrire :
$$\gamma _t=\stackrel{~}{\gamma }_t+t^d\gamma _d.$$
Grรขce ร lโhypothรจse de rรฉcurrence on peut รฉcrire :
$$\begin{array}{cc}\hfill \gamma _t& =e^{t^d\gamma _d}\stackrel{~}{\gamma }_t\hfill \\ & =e^{t^d\gamma _d}e^{tD_1+\mathrm{}+t^rD_r}\gamma _0.\hfill \end{array}$$
Le terme $`\gamma _d`$ appartient ร lโalgรจbre de Lie $`\text{g}^0\text{m}`$. La sรฉrie de Campbell-Hausdorff ne comprend quโun nombre fini de termes dans le cas dโun groupe nilpotent, ce qui permet de conclure. Le passage aux limites projectives dโalgรจbres commutatives nilpotentes de dimension finie se fait sans difficultรฉ.
$``$
Compte tenu de la proposition $`V.2`$, la proposition A.2.1 entraรฎne le rรฉsultat suivant :
Thรฉorรจme A.2.2.
Soient $`\text{g}_1`$ et $`\text{g}_2`$ deux $`L_{\mathrm{}}`$-algรจbres quasi-isomorphes. Alors les foncteurs de dรฉformation de $`\text{g}_1`$ et de $`\text{g}_2`$ sont isomorphes.
En particulier les classes dโรฉquivalence de jauge de $`2`$-tenseurs de Poisson formels sur une variรฉtรฉ sont en bijection avec les classes dโรฉquivalence de jauge dโรฉtoile-produits.
Rรฉfรฉrences : \[AKSZ\] M. Alexandrov, M. Kontsevich, A. Schwarz, O. Zaboronsky, The geometry of the Master equation and topological quantum field theory, Int. J. Mod. Phys. et hep-th 9502010. \[BFFLS\] F. Bayen, M. Flato, C. Frรธnsdal, A. Lichnerowicz, D. Sternheimer, Deformation theory and quantization I. Deformations of symplectic structures, Ann. Phys. 111 No1, 61-110 (1978). \[FM\] W. Fulton, R. MacPherson, Compactification of configuration spaces, Ann. Math. 139, 183-225 (1994). \[H-S\] V. Hinich, V. Schechtman, Homotopy Lie algebras, I.M. Gelfand Seminar, Adv. Sov. math. 16 (2), 1993. \[Kh\] A.G. Khovanskiฤญ, On a lemma of Kontsevich. Funct. Anal. Appl. 31, no. 4, 296โ298 (1998) \[K1\] M. Kontsevich, Deformation quantization of Poisson manifolds, q-alg. 9709040. \[K2\] M. Kontsevich, Formality conjecture, D. Sternheimer et al. (eds.), Deformation theory and symplectic geometry, Kluwer, 139-156 (1997). \[M\] S. Majid, Algebras and Hopf algebras in braided categories, q-alg. 9509023. \[Q\] D. Quillen, Rational homotopy theory, Ann. Math. 90 (1969), 205-295. |
warning/0003/hep-th0003292.html | ar5iv | text | # 1 Introduction
## 1 Introduction
In the original paper Nambu introduced his famous Poisson bracket, which describes a generalized Hamilton mechanics. Let $`(x^i,i=1,2,\mathrm{},n)`$ be a set of dynamical variables which span a $`n`$-dimensional phase space. A Nambu-Poisson structure is defined as
$`\{f_1,f_2,\mathrm{},f_n\}:={\displaystyle \underset{i_1,i_2,\mathrm{},i_n}{}}ฯต_{i_1,i_2,\mathrm{},i_n}{\displaystyle \frac{f_1}{x_{i_1}}}{\displaystyle \frac{f_2}{x_{i_2}}}\mathrm{}{\displaystyle \frac{f_n}{x_{i_n}}},`$ (1)
where $`ฯต_{i_1,i_2,\mathrm{},i_n}`$ is the $`n`$-dimensional Levi-Civita tensor. This bracket has the following properties
1. Skew-symmetry :
$`\{f_1,f_2,\mathrm{},f_n\}=(1)^{ฯต(p)}\{f_{p(1)},f_{p(2)},\mathrm{},f_{p(n)}\},`$
where $`p(i)`$ is the permutation of indices and $`ฯต(p)`$ is the parity of the permutation.
2. Derivation :
$`\{f_1f_2,f_3,\mathrm{},f_{n+1}\}=f_1\{f_2,f_3,\mathrm{},f_{n+1}\}+\{f_1,f_3,\mathrm{},f_{n+1}\}f_2.`$
3. Fundamental Identity (FI):
$`\{\{f_1,f_2,\mathrm{},f_n\},f_{n+1},\mathrm{},f_{2n1}\}=\{\{f_1,f_{n+1},\mathrm{},f_{2n1}\},f_2,\mathrm{},f_n\}`$
$`+\{f_1,\{f_2,f_{n+1},\mathrm{},f_{2n1}\},f_3,\mathrm{},f_n\}+\mathrm{}+\{f_1,\mathrm{},f_{n1},\{f_n,f_{n+1},\mathrm{},f_{2n1}\}\},`$
which is a generalization of the Jacobi identity.
The dynamics of a Nambu system is determined by $`(n1)`$ Hamiltonians $`H_1,H_2,\mathrm{},H_{n1}`$ and is described by Nambu-Hamiltonian equation
$`{\displaystyle \frac{df}{dt}}=\{f,H_1,H_2,\mathrm{},H_{n1}\}.`$ (2)
The quantization of Nambu bracket turns out to be a quite non-trivial problem. In general one expects that the quantum Nambu bracket satisfies some properties analogous to those of the classical Nambu bracket, i.e.
1. Skew-symmetry,
$`[A_1,A_2,\mathrm{},A_n]=(1)^{ฯต(p)}[A_{p(1)},A_{p(2)},\mathrm{},A_{p(n)}].`$ (3)
where $`p(i)`$ is the permutation of indices and $`ฯต(p)`$ is the parity of the permutation.
2. Derivation :
$`[A_1A_2,A_3,\mathrm{},A_{n+1}]=A_1[A_2,A_3,\mathrm{},A_{n+1}]+[A_1,A_3,\mathrm{},A_{n+1}]A_2.`$ (4)
3. Fundamental Identity (FI):
$`[[A_1,A_2,\mathrm{},A_n],A_{n+1},\mathrm{},A_{2n1}]=[[A_1,A_{n+1},\mathrm{},A_{2n1}],A_2,\mathrm{},A_n]`$ (5)
$`+[A_1,[A_2,A_{n+1},\mathrm{},A_{2n1}],A_3,\mathrm{},A_n]+\mathrm{}+[A_1,\mathrm{},A_{n1},[A_n,A_{n+1},\mathrm{},A_{2n1}]].`$
To maintain the skew-symmetry, it is natural to introduce the following quantum commutator
$`[A_1,A_2,\mathrm{},A_n]:={\displaystyle \underset{i_1,i_2,\mathrm{},i_n}{}}ฯต_{i_1,i_2,\mathrm{},i_n}(A_{i_1}A_{i_2}\mathrm{}A_{i_n}).`$ (6)
Thus the problem of quantizing Nambu bracket reduces to finding a proper quantum product, which maps $`n`$ objects to one, such that the quantum commutator (6) satisfies the derivation law (4) and the fundamental identity (5). This turns out to be a very difficult problem. Several quantization schemes have been proposed in the literatures. Among them, the Zariski quantization is a deformation quantization, which obeys all the three requirements listed in the above, and has a clear relationship with the classical Nambu bracket . In the other approaches, the quantum Nambu bracket does not satisfy the derivation law . In particular, Awata etc. proposed a matrix model realization of the quantum Nambu bracket, but its link to the classical Nambu bracket is quite obscure. We will give another matrix realization of the quantum Nambu bracket, which redues in the continuum limit to the classical Nambu bracket.
## 2 Quantum Nambu bracket
Let each of $`\{A_i,i=1,2,\mathrm{},n\}`$ denote a $`\stackrel{n}{\stackrel{}{N\times N\times \mathrm{}\times N}}`$ matrix, and $`๐ฎ_n(N)`$ the collection of all such matrices. we define the quantum product as follows <sup>*</sup><sup>*</sup>*Throughout the paper, any summation index will be specified explicitly.
$`(A^{(1)}A^{(2)}\mathrm{}A^{(n)})_{i_1i_2,\mathrm{},i_n}:={\displaystyle \underset{l}{}}A_{i_1i_2\mathrm{}i_{n1}l}^{(1)}\mathrm{}A_{i_1li_3\mathrm{}i_n}^{(n1)}A_{li_2\mathrm{}i_n}^{(n)}.`$ (7)
We may also define the inner scalar product, i.e. the โTrace
$`\mathrm{Tr}(A):={\displaystyle \underset{i=1}{\overset{N}{}}}A_{ii\mathrm{}i}.`$ (8)
When $`n=2`$, we recover the usual product of the square matrix. When $`n3`$, the quantum product (7) is non-commutative and non-associative. The derivation law (4) does not hold since the product of two objects $`A_1`$ and $`A_2`$ does not make sense. Even worse, the above quantum product generally does not make the quantum commutator (6) satisfying the Fundamental Identity (5). We would expect that there exists some subset of $`๐ฎ_n(N)`$, which is closed under the quantum commuator (6), and obeys on the Fundamental Identity (5). We will give some examples to show that this is indeed the case.
In this article we only consider the quantization of the the even-dimensional Nambu bracket. To do so, set $`n=2m(m2)`$. Let $`๐๐ฎ_{2m}(N)`$ denote the subset of $`๐ฎ_{2m}(N)`$ containing all matrices which are totally antisymmetric, i.e.
$`A_{i_1i_2\mathrm{}i_{2m}}=(1)^{ฯต(p)}A_{p(i_1)p(i_2)\mathrm{}p(i_{2m})}.`$
It is obvious to see that this trucation is consistent with the quantum commutator(6), if we use the quantum product (7). When $`N<2m`$, the subset $`๐๐ฎ_n(N)`$ is empty. When $`N=2m`$, $`๐๐ฎ_n(N)`$ contains just one element, which generates an abelian algebra. So the Fundamental Identity is trivial. The first non-trivial example is the case $`N=2m+1`$, where $`๐๐ฎ_{2m}(N)`$ contains $`N`$ independent elements, which we may denote by
$`(T_a)_{i_1i_2\mathrm{}i_{2m}}:=ฯต_{ai_1i_2\mathrm{}i_{2m}},a=1,2,\mathrm{},N.`$ (9)
It is straightforward to show that these matrices generate a quantum algebra with the following commutation relation
$`[T_{a_1},T_{a_2},\mathrm{},T_{a_{2m}}]={\displaystyle \underset{b}{}}ฯต_{a_1a_2\mathrm{}a_{2m}b}T_b.`$ (10)
The fundamental identity is guaranteed by the following equality
$`{\displaystyle \underset{c}{}}(ฯต_{ca_1a_2\mathrm{}a_n}ฯต_{cb_1b_2\mathrm{}b_n}{\displaystyle \underset{i=1}{\overset{n}{}}}(a_ib_n))=0.`$ (11)
One may generalize the consideration here to the case with the general value of $`N`$.
## 3 The continuum limit
The quantum product (7) has a very simple geometric meaning. Consider $`N(Nn)`$ ordered points in $`(n1)`$-dimensional Euclidean space $`๐^{n1}`$. For each $`(n1)`$-dimensional convex with vertices $`(i_1,i_2,\mathrm{},i_n)`$ in general positions, we associate to it a matrix element $`A_{i_1,i_2,\mathrm{},i_n}`$. Adding a point $`l`$ to this convex leads to $`n`$ more convexes, all of which contain the point $`l`$. Thus we have a natural composition law,
$`\underset{n}{\underset{}{๐ฎ\times ๐ฎ\times \mathrm{}\times ๐ฎ}}๐ฎ,(A_{i_1i_2,\mathrm{},i_n})๐ฎ,`$ (12)
which glues $`n`$ convexes to one, and defines a general โgroup structureโ. The explicit form of this composition law is given by the quantum product (7).
The geometric picture provides a way to analyse the continuum limit of the quantum Nambu bracket. To do so, we may take the following viewpoint. Each matrix defines a map from the set of convexes to the field $`๐`$ or $`๐`$,
$`A:\{\mathrm{convexes}\}๐(\mathrm{or}๐).`$
Each convex can be characterized by its center and volume We may think of that the ordered points are sited on a $`(n1)`$-dimensional rectangular lattice. The volume-preserving symmetry group of the lattice is $`sl(n1,๐)`$. We are interested in the quotient lattice space, which is the regular lattice modulo out $`sl(n1,๐)`$ symmetry., which is
$`\stackrel{}{x}:={\displaystyle \frac{ฯต}{n}}(\stackrel{}{i}_1+\stackrel{}{i}_2+\mathrm{}+\stackrel{}{i}_n)=(x_1,x_2,\mathrm{},x_{n1}),ฯต:={\displaystyle \frac{1}{N}},`$
$`\mathrm{vol}:={\displaystyle \frac{1}{(n1)!}}ฯต_{\mu _1\mu _2\mathrm{}\mu _{n1}}(\stackrel{}{i}_1\stackrel{}{i}_n)_{\mu _1}(\stackrel{}{i}_2\stackrel{}{i}_n)_{\mu _2}\mathrm{}(\stackrel{}{i}_{n1}\stackrel{}{i}_n)_{\mu _{n1}},`$
$`\stackrel{}{v}^{(j)}=\stackrel{}{i}_{nj+1}\stackrel{}{l},1jn.`$
Where $`(\mu _1,\mu _2,\mathrm{},\mu _{n1})`$ are indices of the Cartesian coordinates in the $`(n1)`$-dimensional space. To go to the continuum limit, we let all the $`n`$-indices $`(i_1i_2,\mathrm{},i_n)`$ go to infinity, but their differences finite. Thus $`\stackrel{}{x}`$ become the continuum variables. To characterize the dependence of a matrix on the volume of the convex, we introduce a new parameter $`x_n`$, such that
$`{\displaystyle \frac{f(\stackrel{}{x},x_n)}{x_n}}:=f(\stackrel{}{x},x_n)\mathrm{vol},`$
here $`f(\stackrel{}{x},x_n)`$ is the continuum limit of the matrix $`A_{i_1i_2,\mathrm{},i_n}`$. One may think of $`x_n`$ as the constant density of mass, momentum or charge, etc. With these notations, we may expand the commutator (6) in the powers of $`ฯต`$, and we find that
$`[A^{(1)},A^{(2)},\mathrm{},A^{(n)}]=({\displaystyle \frac{n!ฯต^{n1}}{n^{n1}}})\{f^{(1)},f^{(2)},\mathrm{},f^{(n)}\}+๐ช(ฯต^n).`$ (13)
## 4 Conclusion
In this letter we have constructed a quantum Nambu bracket via matrices of multi-index. We have shown that our quantum commutator has the right continuum limit. Our construction may be helpful to quantize $`p`$-brane theory, just like what we did in studying $`2`$-dimensional quantum gravity by the ordinary matrix models. Furthermore if we regard the points in $`๐^{n1}`$ as a set of $`D0`$-particles, we may expect a matrix theory in more general setting. We wish to discuss these problems in more details elsewhere.
Acknowledgements
We thank Drs. H. Awata, M. Li, X.C. Song, M. YU and C. J. Zhu for discussions and comments. The work is supported in part by NSFC grant 19925521 and by the Startup grant from Beijing University. |
warning/0003/quant-ph0003125.html | ar5iv | text | # Trapped ions in laser fields: a benchmark for deformed quantum oscillators
## I Introduction
The theory of certain one-parameter (qโ) deformations of Lie algebras, the so called quantum groups has been of great interest in the last decade in several areas of physics. In 1989 Biedenharn and McFarlane independently defined the q-analogue coherent state of a deformed q-oscillator, for which Nelson et al. were able to obtain the resolution of unity. Since then the properties of a class of deformations of the harmonic oscillator were considered by several authors (see f.i. ). Deformed quantum oscillators are represented by dynamical variables $`A`$, $`A^{}`$ and $`N_A`$ satisfying the commutation relations, $`[A,N_A]=A,`$ $`[A^{},N_A]=A^{}`$ and $`[A,A^{}]=f\left(N_A\right)`$, with $`f\left(N_A\right)`$ an arbitrary real function of $`N_A`$. All such variables are constructed in terms of single-mode field operators<sup>*</sup><sup>*</sup>*wherever possible operators will be indicated by simple letters, except for the addition of a caret when confusion could arise with c-number quantities $`a,a^{}`$ and $`a^{}a.`$
In 1993 Crnugelj et al. observed that the multiphoton interaction of a single mode laser field with a two level atom is described by deformed-oscillator creation and annihilation operators which in combination with the pseudo-spin atomic operators $`\sigma _+`$ and $`\sigma _{}`$, form the potential
$`W_{JC}=A^{}\sigma _{}+A\sigma _+`$
used in the Jaynes-Cummings model (JCM) .
In that period the J-C model was at center of the attention for the study of laser cooling of ions placed in parabolic traps, with the quantized center-of-mass motion of the ion playing the role of the boson mode, coupled via the laser to the internal degrees of freedom. In the case of cooling the operator $`A`$ is represented by a combination of some power of the annihilation operator $`a`$ times a function of $`n`$. When some bosons of the oscillator mode are destroyed the ion is excited to the upper level from where it decays radiatively. Cooling was investigated in Lamb-Dicke and strong-sideband limits, that is for ion excursions small compared with the radiation wavelength. This study led to the discovery of many intriguing effects connected with the nonclassical properties of the field, like as a long-time sensitivity to the statistical properties of the radiation field . For example, the mean excitation number of the quantized oscillations of a ion driven by a squeezed field exhibited periodic collapses and revivals .
The interest for the vibrational motion of trapped ions was also motivated by the connection between the state of motion and the properties of the fluorescence spectra . This link led some experimentalists to look for new non classical radiation states generated by trapped ions forced into some unusual vibrational states. In analogy to the preparation of nonclassical states of light in quantum optics several authors examined the preparation of the center-of-mass motion in a quantum state having no classical counterpart. Worthy examples were those of Cirac et al. who considered the possibility of generating squeezed states of the vibrational motion by irradiating the trapped ion with two standing-wave light fields of different frequencies and locating the center of the trap potential at a common node of both waves. In all these cases the nonlinear dependence of $`A`$ on $`a,a^{}`$ and $`\widehat{n}`$, stemmed from the ion motion in the trap potential.
de Matos Filho and Vogel observed in 1993 that the center-of-mass state of a trapped ion driven by a two-mode laser field decays toward a dark state coincident with a nonlinear coherent state (hereinafter called NCS) of a deformed oscillator. This result brought new fuel to the study of deformed oscillators describing different classes of states arising in the trapped ion motion under the action of two or three fields detuned by multiples of the vibrational frequency (see f.i. for nonlinear cat states). In the wake of this interest attention was paid to theoretical models of deformed oscillators, like those connected with excited coherent states and binomial states . All these non-linear oscillators differ for the deformation function $`h\left(\widehat{n}\right)`$ connecting the annihilation operator $`a`$ to the deformed one $`A=ah\left(\widehat{n}\right)`$. The ancestor of these realizations were the q-oscillators characterized by a deformation $`h_q\left(\widehat{n}\right)=\sqrt{\frac{\mathrm{sinh}\left(\lambda \widehat{n}\right)}{\widehat{n}\mathrm{sinh}\lambda }}`$ increasing with $`n`$. Contrarily the trapped ion deformation is a very irregular function of $`n`$, taking positive and negative values. What is worse, for some combinations of the Lamb-Dicke parameter $`\eta ^2`$ and $`n`$ it can vanish or become infinite. As a consequence it is hard to capitalize on the work done for the q-oscillator for studying the NCS of a trapped ion. In particular, while for the q-case it has been found a measure resolving the unity, the same is not exactly true for the ion case. As a consequence the formalism of the Bargmann spaces , which has been extended from the linear oscillators to the q-ones, cannot be applied exactly to the ion case. In fact, it will be shown in the following that this can be done by considering a class of rational deformations which approximate to any degree of accuracy the ion deformation. In most experimental cases the statistical state of a trapped ion is limited to a finite number of Fock states so that these rational deformations may adequately approximate the ion deformation. Only in this โweakโ sense it is possible to construct a ion-analogue of a Bargmann space, on which the deformed creation and annihilation operators are represented as multiplication by $`z`$ and differentiation with respect to $`z`$, respectively.
This paper is dedicated to an extension of the theory of the usual coherent states to NCS using as examples the deformation relative to the dark states of trapped ions. We start with a single-mode excitation field $`A`$ (Sec. II), by discussing some properties of NCS, and introducing a deformed version $`D_h\left(\alpha \right)`$ of the displacement operator (Sec. III). In Sec. IV we discuss some aspects of the resolution of unity for these NCS. The operator $`D_h\left(\alpha \right)`$ is used in Sec. V for associating the density matrix operator $`\widehat{\rho }`$ to a linear functional $`\rho _{A,h}\left(z\right)`$ mapping the test function $`\mathrm{exp}\left(\alpha z^{}\alpha ^{}z\right)`$ into the expectation value $`D_h\left(\alpha \right)`$, by extending the construction of the antinormal probability distribution function . The connection with the P-representation is also briefly examined. Section VI is dedicated to NCS on a circle, for which the Wigner functions are presented. Finally, the last section is dedicated to the dark states, arising when a trapped ion is driven by a bichromatic laser field. An asymptotic expression of the deformation and the relative factorial is obtained and its implication on the convergence of the NCS series is discussed. It comes out that it converges only for $`\alpha `$ in a circle of radius equal to the inverse of $`\eta `$. On the other hand the weight of each Fock state can take values so large to prevent the resolution of unity in terms of normalized NCS. Some approximate expressions of the deformation are discussed together with the possibility of using these NCS for representing the ion statistical state.
## II Motion of a trapped and laser-driven ion
We consider an ideal two-level ion of mass M constrained to move in a 3D harmonic potential. Taking the principal trap (x-axis) axis to coincide with the direction of propagation of the driving field, one quantum number suffices to label the vibrational states of the trap. The other two are traced out by summing over the corresponding degrees of freedom.
The ionโs internal and external degrees of freedom are coupled together by a light field $`e^{i\omega _Lt+i\phi \left(t\right)}`$ periodically modulated at the frequency $`\nu `$ of the ion trap
$`E(x,t)=e^{i\omega _Lt+i\phi \left(t\right)}g\left(t\right)f\left(x\right)+h.c.`$
where $`g\left(t\right)=g\left(t+\frac{2\pi }{\nu }\right)`$ is a generally complex periodic function of frequency $`\nu `$ and $`h.c.`$ stays for the Hermitian conjugate. The function $`f\left(x\right)`$ stands for $`e^{ik_Lx}`$ or $`\mathrm{sin}\left(k_Lx+\varphi \right)`$ respectively for a progressive or standing wave, with the phase $`\varphi `$ determining the position of the trap potential with respect to the standing wave.
We will dwell on monochromatic
$`g_{N+1}^{\left(1\right)}\left(t\right)=e^{i\left(N+1\right)\nu t}`$
and bichromatic driving fields
$`g_{N+1}^{\left(2\right)}\left(t\right)=e^{i\left(N+1\right)\nu t}\alpha _{N+1}`$
with the parameter $`N`$ taking non-negative integer values, and $`\alpha _{N+1}`$ a complex coefficient depending on the amplitudes of the two waves.
Now, introducing the Lamb-Dicke parameter $`\eta =\mathrm{}k_L/\sqrt{2M\mathrm{}\nu }`$ we put as usual $`e^{ikx}=e^{i\eta \left(a_v^{}+a_v\right)}`$. In the classical limit $`\eta `$ is large and the absorption or emission of a photon will always cause some change in the vibrational state of the atom. In the non-classical Lamb-Dicke limit (LDL) of small $`\eta `$, many photons may need to be absorbed or emitted before the atom changes vibrational state. For example in the sideband cooling experiment carried out by Diedrich et al. the parameter $`\eta `$ was equal to $`0.06`$.
The Hamiltonian for a trapped ion interacting with a bichromatic field can be split in two parts
$`H=H_0+H_{int}`$
where ($`\mathrm{}=1`$)
$$H_0=\omega _{12}\sigma _3+\nu \widehat{n}$$
(1)
and, in the electric dipole approximation,
$`H_{int}=\mathrm{}\left(\sigma _{}E^{}(x,t)+\sigma _+E(x,t)\right)`$
When the Rabi frequency $`\mathrm{\Omega }`$, relative to the laser induced transition between the ion ground and excited levels, is much smaller than the trapping potential frequency $`\nu `$, a perturbation expansion can be carried out in $`\mathrm{\Omega }/\nu `$, as discussed in Ref. . This expansion allows a division into quickly and slowly varying density operator matrix elements, the former of which can be adiabatically eliminated.
Arresting the calculation to the zeroth-order in $`\mathrm{\Omega }/\nu `$ amounts to applying the rotating wave approximation. This approach can be easily pursued by switching to the interaction picture defined by the unitary operator $`U_{rw}=\mathrm{exp}\left[i\left(\omega _L\sigma _3+\nu \widehat{n}\right)t\right]`$ and retaining in the transformed hamiltonian $`H^{}`$ the time-independent terms together with the slowly varying phase $`\phi \left(t\right)`$ of the laser field,
$$H^{}=\left(\mathrm{\Delta }\dot{\phi }\left(t\right)\right)\sigma _3+\mathrm{\Omega }\left(\sigma _{}A+\sigma _+A^{}\right)$$
(2)
with $`\mathrm{\Delta }=\omega _{12}\omega _L`$ the detuning parameter, $`\mathrm{\Omega }=e^{\eta ^2/2}\mathrm{}`$ the vibronic Rabi frequency and
$$A=e^{\eta ^2/2}\overline{g\left(t\right)f\left[\eta \left(e^{i\nu t}a^{}+e^{i\nu t}a\right)\right]}$$
(3)
the bar indicating the time average.
Expanding the factor $`e^{i\eta \left(e^{i\nu t}a^{}+e^{i\nu t}a\right)}`$ in power series in $`a`$ and $`a^{}`$, introducing the operator
$$f_k(\widehat{n},\eta ^2)=\underset{m=0}{\overset{\mathrm{}}{}}\frac{\left(\widehat{n}m+1\right)_m}{\left(k+1\right)_mm!}\left(\eta ^2\right)^m=k!\frac{L_{\widehat{n}}^k\left(\eta ^2\right)}{\left(\widehat{n}+1\right)_k}$$
(4)
with $`\left(\widehat{n}m+1\right)_m=\left(a^{}\right)^ma^m=\widehat{n}\left(\widehat{n}1\right)\mathrm{}\left(\widehat{n}m+1\right)`$ and $`L_{\widehat{n}}^k\left(\eta ^2\right)`$ reducing in the Fock basis to the generalized Laguerre polynomials, we obtain respectively for progressive
$$e^{i\eta \left(e^{i\nu t}a^{}+e^{i\nu t}a\right)}=e^{\eta ^2/2}\underset{k=0}{\overset{\mathrm{}}{}}ฯต_k\frac{\left(i\eta \right)^k}{k!}\left[f_k\left(\widehat{n}\right)a^ke^{ik\nu t}+\left(a^{}\right)^kf_k\left(\widehat{n}\right)e^{ik\nu t}\right]$$
(5)
and standing waves
$$\mathrm{sin}\left[\eta \left(a^{}+a\right)+\varphi \right]=e^{\eta ^2/2}\underset{k=0}{\overset{\mathrm{}}{}}ฯต_k\frac{\left(\eta \right)^k}{k!}\mathrm{sin}\left(\varphi +k\frac{\pi }{2}\right)\left[a^kf_k\left(\widehat{n}+k\right)e^{ik\nu t}+f_k\left(\widehat{n}+k\right)\left(a^{}\right)^ke^{ik\nu t}\right]$$
(6)
with $`k`$ a positive integer and $`ฯต_k=\frac{1}{2}`$ for $`k=0`$ and $`ฯต_k=1`$ otherwise.
For progressive $`\left(p\right)`$ and stationary $`\left(s\right)`$ monochromatic waves with $`g\left(t\right)=e^{i\left(N+1\right)\nu t}`$ the operator $`A`$ (see (3)) is given by
$`A_p^{\left(1\right)}`$ $`=`$ $`{\displaystyle \frac{\left(i\eta \right)^{N+1}}{\left(N+1\right)!}}f_{N+1}(\widehat{n},\eta ^2)a^{N+1}`$ (7)
$`A_s^{\left(1\right)}`$ $`=`$ $`\left(i\right)^{N+1}\mathrm{sin}\left(\varphi +(N+1){\displaystyle \frac{\pi }{2}}\right)A_p^{\left(1\right)}`$ (8)
while for two modes bichromatic driving fields
$`A_p^{\left(2\right)}`$ $`=`$ $`f_{N+1}\left(\widehat{n}\right)a^{N+1}\alpha _{N+1}f_0\left(\widehat{n}\right)`$ (9)
$`A_s^{\left(2\right)}`$ $`=`$ $`\left(i\right)^{N+1}\mathrm{sin}\left(\varphi +(N+1){\displaystyle \frac{\pi }{2}}\right)f_{N+1}\left(\widehat{n}\right)a^{N+1}+\mathrm{sin}\left(\varphi \right)\alpha _{N+1}f_0\left(\widehat{n}\right).`$ (10)
## III Nonlinear coherent states
Coherent states were originally introduced as eigenstates of the annihilation operator for the harmonic oscillator . They have been generalized (see ) by labeling as nonlinear coherent states $`|\alpha ,h`$ the right-hand eigenstates
$$A|\alpha ,h=\alpha |\alpha ,h$$
(11)
of operatorsfor the sake of notational simplicity we will use the same symbol $`A`$ for indicating fields of the form (8) and (10). $`A`$ of the form
$$A=ah\left(\widehat{n}\right)$$
(12)
where $`h\left(\widehat{n}\right)`$ is an operator-valued real function of the number operator. It is immediate to show that
$$|\alpha ,h=N_{h,\alpha }\underset{n=0}{\overset{\mathrm{}}{}}\frac{\alpha ^n}{\sqrt{n!}\left[h\left(n\right)\right]!}|n$$
(13)
with $`\left[h\left(n\right)\right]!=h\left(0\right)h\left(1\right)\mathrm{}h\left(n\right)`$ and normalizing factor $`N_{h,\alpha }`$
$`N_{\alpha ,h}={\displaystyle \frac{1}{\sqrt{E_h\left(\left|\alpha \right|^2\right)}}}`$
expressed in terms of the entire function
$`E_h\left(v\right)={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{v^n}{n!\left(\left[h\left(n\right)\right]!\right)^2}}`$
referred to in the following as h-exponential in analogy with the q-exponential used in Ref. .
The deformation functions $`h(n,\eta ^2)`$ associated to the dark states of the trapped ions are represented by the ratio of two Laguerre polynomials of argument equal to the LambโDicke parameter $`\eta ^2`$ so that they vanish or become infinite for some isolated combinations of $`\eta ^2`$ and $`n`$. We are obliged to explicitly assume that this situation does not occur for the values of $`\eta ^2`$ considered.
In Sec. VII we will obtain an asymptotic expression of the weights of the Fock states occurring in the series expansion of the NCS relative to trapped ions. They take very large and very small values for increasing $`n`$, so that these NCS can be normalized only for $`\alpha `$ inside the circle $`1/\eta .`$ For convenience of discussion we shall ignore this problem by restricting our treatment here to normalized NCS states.
It may be worth noting at this point that many of the foregoing formulas may be abbreviated by adopting a normalization different from the conventional one for the coherent state. If we introduce the symbol $`\alpha ;h`$ for the states normalized in the new way and define these as
$`\alpha ;h=N_{\alpha ,h}^1|\alpha ;h`$
and
$`\alpha ;h\beta ;h=E_h\left(\alpha ^{}\beta \right).`$
Since the commutator $`[A^{},A]=\widehat{n}h^2\left(\widehat{n}\right)(\widehat{n}+1)h^2(\widehat{n}+1)`$ is not a c-number it is worthy introducing the operator
$$A_h^{}=\frac{1}{h\left(\widehat{n}\right)}a^{}$$
(14)
With these alterations we have
$`A\alpha ,h`$ $`=`$ $`\alpha \alpha ,h,`$
$`\widehat{n}\alpha ,h`$ $`=`$ $`\alpha {\displaystyle \frac{}{\alpha }}\alpha ,h,`$
$`A_h^{}\alpha ,h`$ $`=`$ $`{\displaystyle \frac{}{\alpha }}\alpha ,h.`$
In addition
$$A^{}\alpha ,h=h^2\left(\widehat{n}\right)A_h^{}\alpha ,h=h^2\left(\alpha \frac{}{\alpha }\right)\frac{}{\alpha }\alpha ,h$$
(15)
In all the above r.h.s. the operators $`\alpha ,_\alpha `$ and their combination are intended to act on the coefficients of the Fock states series.
### A Displacement and deformation operators
It is well known that coherent state $`|\alpha `$ can be also introduced by displacing the Fock vacuum state $`|0`$ by means of the operator
$$๐\left(\alpha \right)=\mathrm{exp}\left(\alpha ^{}a+\alpha a^{}\right)$$
(16)
due to its property of displacing the annihilation operator $`a`$ by the generally complex quantity $`\alpha `$,
$`๐\left(\alpha \right)a๐\left(\alpha \right)=a\alpha `$
Unfortunately, $`๐\left(\alpha \right)`$ is unable to displace the deformed operator $`A`$. In alternative $`๐\left(\alpha \right)`$ could be replaced by the unitary operator obtained by replacing in Eq. (16) $`a`$ and $`a^{}`$ by $`A`$ and $`A^{}`$ respectively, but also this operator does not displace $`A`$ by the complex quantity $`\alpha `$. The difficulties in dealing with exponentials of linear combinations of $`A`$ and $`A^{}`$ originate from the circumstance that their commutator is not a c-number. These problems can be overcome by using $`A_h^{}`$ (see Eq. (14)) in place of $`A^{}`$ and defining the โdeformedโ version of the displacement operator as
$$๐_h\left(\alpha \right)=\mathrm{exp}\left(\alpha ^{}A+\alpha A_h^{}\right)=e^{\frac{\left|\alpha \right|^2}{2}}e^{\alpha ^{}A}e^{\alpha A_h^{}}=e^{\frac{\left|\alpha \right|^2}{2}}e^{\alpha A_h^{}}e^{\alpha ^{}A}$$
(17)
$`๐_h\left(\alpha \right)`$ shares many properties of the standard operator $`๐\left(\alpha \right)`$ as
$`๐_h^1\left(\alpha \right)=๐_h\left(\alpha \right)`$
and
$`๐_h\left(\beta \right)๐_h\left(\alpha \right)=\mathrm{exp}\left[{\displaystyle \frac{1}{2}}\left(\beta \alpha ^{}\beta ^{}\alpha \right)\right]๐_h\left(\beta +\alpha \right)`$
However, $`๐_h\left(\alpha \right)`$ is not a unitary operator,
$`๐_h^{}\left(\alpha \right)=๐_{\frac{1}{h}}^1\left(\alpha \right)=๐_{\frac{1}{h}}\left(\alpha \right)`$
so that it does not preserve the norm of a state.
$`๐_h\left(\alpha \right)`$ and $`๐_{\frac{1}{h}}\left(\alpha \right)`$ displace $`A`$ and $`A^{}`$ respectively by $`\alpha `$ and $`\alpha ^{}`$,
$`๐_h\left(\alpha \right)A๐_h\left(\alpha \right)`$ $`=`$ $`A\alpha `$ (18)
$`๐_{\frac{1}{h}}\left(\alpha \right)A^{}๐_{\frac{1}{h}}\left(\alpha \right)`$ $`=`$ $`A^{}\alpha ^{}`$ (19)
and $`A_h^{}`$ by $`\alpha ^{}`$
$$๐_h\left(\alpha \right)A_h^{}๐_h\left(\alpha \right)=A_h^{}\alpha ^{}$$
(20)
Accordingly, the NCS $`|\alpha ;h`$ can be obtained by applying $`๐_h\left(\alpha \right)`$ to the vacuum state,
$`\alpha ;h=e^{\alpha A_h^{}}|0=e^{\frac{\left|\alpha \right|^2}{2}}๐_h\left(\alpha \right)|0`$
In conclusion, the NCS $`\alpha ;h`$ can be obtained by deforming the usual coherent state $`\alpha `$ by means of the deformation operator
$$d_h=๐_h\left(\alpha \right)๐\left(\alpha \right)$$
(21)
namely,
$`\alpha ;h=d_h\alpha `$
Although expressed as a product of operators depending on the complex parameter $`\alpha ,`$ $`d_h`$ is independent of $`\alpha `$. In a Fock basis it is diagonal with components equal to $`\left[h\left(n\right)\right]!^1`$. Since $`h\left(n\right)`$ does not vanish, as already assumed, $`d_h`$ is not singular.
Finally, we note that
$`m\left|๐_h\left(\alpha \right)\right|n={\displaystyle \frac{\left[h\left(m\right)\right]!}{\left[h\left(n\right)\right]!}}m\left|๐\left(\alpha \right)\right|n`$
so that the matrix representation of $`๐`$ and $`๐_h`$ have the same diagonal part.
A further remark is that the set of operators $`๐_h(\alpha )`$ constitutes a Weyl system which does not lead to the canonical quantization for not being unitary.
### B Nonlinear displaced Fock states
In Sec. V we will use the Fock states displaced by $`๐_h\left(\alpha \right)`$ (see Eq. (30))
$$|\phi _m,\alpha ,h=๐_h\left(\alpha \right)|m$$
(22)
which can be shown with the help of Eqs. (19) and (20) to be the right eigenstates of the operator $`\left(A_h^{}\alpha ^{}\right)\left(A\alpha \right)=๐_h\left(\alpha \right)\widehat{n}๐_h\left(\alpha \right)`$,
$`\left(A_h^{}\alpha ^{}\right)\left(A\alpha \right)|\phi _m,\alpha ,h=m|\phi _m,\alpha ,h`$
Analogously we can introduce the left eigenstates defined by
$`\psi _m,\alpha ,h|\left(A_h^{}\alpha ^{}\right)\left(A\alpha \right)=m\psi _m,\alpha ,h|`$
which are obtained by displacing $`m|`$ by $`๐_h\left(\alpha \right),`$ i.e. $`\psi _m,\alpha ,h|=m|๐_h\left(\alpha \right).`$
It is noteworthy that the left and right displaced Fock states are mutually orthogonal,
$`\psi _m,\alpha ,h|\phi _n,\alpha ,h=0`$
for $`mn.`$
On the other hand these states can be also expressed in the form
$`|\phi _m,\alpha ,h`$ $`=`$ $`{\displaystyle \frac{\left[h\left(m\right)\right]!\left(A_h^{}\alpha ^{}\right)^m}{\sqrt{m!}}}|\alpha ,h={\displaystyle \frac{\left[h\left(m\right)\right]!}{\sqrt{m!}}}{\displaystyle \underset{n}{}}\left({\displaystyle \genfrac{}{}{0pt}{}{m}{n}}\right)\left(\alpha ^{}\right)^{mn}|\alpha ,h,m`$
$`\psi _m,\alpha ,h|`$ $`=`$ $`m|D_h\left(\alpha \right)=\alpha ,h|{\displaystyle \frac{\left(A\alpha \right)^m}{\left[h\left(m\right)\right]!\sqrt{m!}}}={\displaystyle \frac{1}{\left[h\left(m\right)\right]!\sqrt{m!}}}{\displaystyle \underset{n}{}}\left({\displaystyle \genfrac{}{}{0pt}{}{m}{n}}\right)\left(\alpha \right)^{mn}\alpha ,h,m|`$
where $`|\alpha ,h,m=A_h^m|\alpha ,h`$ and $`\alpha ,h,m|=\alpha ,h|A^m`$ stand for the deformed versions of the excited coherent states (see also ).
## IV Resolution of the unity
From the completeness relation of coherent states
$`1={\displaystyle \frac{1}{\pi }}{\displaystyle |\alpha \alpha |d^2\alpha }`$
it descends
$`1={\displaystyle \frac{1}{\pi }}{\displaystyle d_h|\alpha \alpha \left|d_h^1d^2\alpha =\frac{1}{\pi }d_h^1\right|\alpha \alpha |d_hd^2\alpha }`$
Next, using the relation
$`d_h^1=๐\left(\alpha \right)๐_{\frac{1}{h}}^{}\left(\alpha \right)=\left(๐_{\frac{1}{h}}\left(\alpha \right)๐\left(\alpha \right)\right)^{}=d_{\frac{1}{h}}^{}`$
the above resolution of unity can be expressed in terms of deformed coherent states
$$1=\frac{1}{\pi }\frac{e^{\alpha \alpha }}{N_{\alpha ,h}N_{\alpha ,\frac{1}{h}}}|\alpha ,h\alpha ,h^1\left|d^2\alpha =\frac{1}{\pi }\frac{e^{\alpha \alpha }}{N_{\alpha ,h}N_{\alpha ,\frac{1}{h}}}\right|\alpha ,h^1\alpha ,h|d^2\alpha $$
(23)
It goes without saying that this resolution holds true only if the NCS relative to the deformations $`h`$ and $`1/h`$ are both normalizable in the whole complex $`\alpha `$โplane.
For some deformations anyhow it is possible to obtain a resolution of unity in terms of projectors of deformed coherent states, i.e. to find a suitable element of measure $`d\mu `$ such that
$$1=\alpha ,h\alpha ,h๐\mu $$
(24)
$`d\mu `$ can be considered as an extension of the measure element $`d\mu =\frac{1}{\pi }e^{\left|\alpha \right|^2}d^2\alpha `$ for the linear oscillators. Since $`m\alpha ,h\alpha ,hn๐\mu `$ must vanish for $`mn`$ $`d\mu `$ can be put in the form
$`d\mu ={\displaystyle \frac{1}{\pi }}m_h\left(\left|\alpha \right|^2\right)d^2\alpha `$
where $`m_h\left(x\right)`$ is a distribution satisfying the set of equations
$$n!\left(\left[h\left(n\right)\right]!\right)^2=m_h\left(x\right)x^n๐x$$
(25)
for every integer $`n`$.
Treating $`n=s1`$ as a continuous variable the above relation represents a Mellin integral transform,
$$g\left(s\right)=_0^{\mathrm{}}f\left(x\right)x^{s1}๐x$$
(26)
so that $`m_h\left(x\right)`$ is the Mellin antitransform of $`g\left(s\right)=\mathrm{\Gamma }\left(s\right)\left(\left[h\left(s1\right)\right]!\right)^2`$.
From the relation $`\beta ,h\left|A^m\right|\beta ,h=\beta ^m`$ it descends that $`E_h\left(\beta ^{}\alpha \right)`$ is the self-reproducing kernel of the h-analogue of the Bargmann space , with respect to $`d\mu `$
$`{\displaystyle \left|E_h\left(\beta ^{}\alpha \right)\right|^2\alpha ^m๐\mu }=\beta ^m`$
In preparation of the discussion of Sec. VII it is worth remarking that replacing $`h`$ by the deformation $`\beta h`$ the relative measure $`m_{\beta h}\left(x\right)`$ is given by
$$m_{\beta h}\left(x\right)=\beta ^2m_h\left(\beta ^2x\right)$$
(27)
This relation can be also used for expressing a thermal density matrix characterized by Boltzmann weight factors $`\rho _{nn}\mathrm{exp}\left(\beta n\right)`$ in the form
$`\widehat{\rho }=e^{2\beta }{\displaystyle \frac{m_h\left(e^{2\beta }\left|\alpha \right|^2\right)}{m_h\left(\left|\alpha \right|^2\right)}\alpha ,h\alpha ,h๐\mu }`$
In Ref. it was possible to obtain the resolution of unity for a q-oscillator by deforming both the derivative and the integral operators while a resolution for the so-called harmonious states was obtained in . We will see in the following that for the trapped ion deformation the measure is a distributional Laplace antitransform which includes non-normalizable NCS.
For a deformation approximated by a rational function of $`n`$, $`g\left(s\right)`$ corresponds to the ratio of products of gamma functions,
$$g\left(s\right)=\frac{\mathrm{\Gamma }\left(a_1+s\right)\mathrm{}\mathrm{\Gamma }\left(a_A+s\right)}{\mathrm{\Gamma }\left(b_1+s\right)\mathrm{}\mathrm{\Gamma }\left(b_B+s\right)}\mathrm{\Gamma }\left[\genfrac{}{}{0pt}{}{\left(a\right)+s}{\left(b\right)+s}\right]$$
(28)
For $`AB`$ the relative antitransform is given by a combination of generalized hypergeometric functions
$`{}_{C}{}^{}F_{D}^{}[{\displaystyle \genfrac{}{}{0pt}{}{\left(c\right)}{\left(d\right)}};\left(1\right)^{C+D+1}x]={\displaystyle \underset{n}{}}{\displaystyle \frac{\left(c_1\right)_n\mathrm{}\left(c_C\right)_n}{\left(d_1\right)_n\mathrm{}\left(d_D\right)_n}}{\displaystyle \frac{x^n}{n!}}\left(1\right)^{n\left(C+D+1\right)}`$
namely
$$m_h\left(x\right)=\underset{\mu =1}{\overset{A}{}}\mathrm{\Gamma }\left[\genfrac{}{}{0pt}{}{\left(a\right)^{}a_\mu }{\left(b\right)a_\mu }\right]_BF_{A1}[\genfrac{}{}{0pt}{}{1+a_\mu \left(b\right)}{1+a_\mu \left(a\right)^{}};\left(1\right)^{A+B}x]x^{a_\mu }$$
(29)
where $`\left(a\right)^{}a_\mu `$ and $`\left(a\right)^{}\left(a\right)^{}a_\mu 1`$ stand for the sequences $`a_1a_\mu ,\mathrm{},a_Aa_\mu ,`$ and $`a_1a_\mu 1,\mathrm{},a_Aa_\mu 1`$ with the exclusion the $`\mu `$-th term.
## V Expansion of statistical states
The same reasons that led to express arbitrary states and operators in term of coherent states, suggest that we develop expansions in terms of NCS as well.
Following we introduce for a statistical state the deformed quantum linear functional
$$F_h\left[\alpha \right]=Tr\left\{\widehat{\rho }๐_h\left(\alpha \right)\right\}=\underset{m}{}\rho _{mn}n|\phi _m,\alpha ,h$$
(30)
In particular for a diagonal density matrix $`F_h\left[\alpha \right]`$ reduces to the standard $`F\left[\alpha \right].`$
Using the unity resolution (23) $`F_h\left[\alpha \right]`$ may be rewritten as
$$F_h\left[\alpha \right]=e^{\frac{\alpha \alpha }{2}}\mathrm{exp}\left(\alpha z^{}\alpha ^{}z\right)\rho _{h,A}\left(z\right)d^2z$$
(31)
where
$`\rho _{h,A}\left(z\right)`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle \frac{e^{zz^{}}}{N_{z,h}N_{z,h^1}}}Tr\left\{\widehat{\rho }|z,hz,h^1|\right\}`$ (32)
$`=`$ $`{\displaystyle \frac{1}{\pi }}z\left|\widehat{\rho }_h\right|z`$ (33)
with
$`\widehat{\rho }_h=d_h^1\widehat{\rho }d_h`$
the deformed density operator. In other words $`\rho _{h,A}\left(z\right)`$ stands for the generalized distribution function of the deformed density matrix $`d_h^1\widehat{\rho }d_h`$.
For extending the definition of the characteristic functional $`F\left[\alpha \right]=Tr\left\{\widehat{\rho }D\left(\alpha \right)\right\}`$ to a deformed oscillator we pay the penalty of loosing some properties of $`\rho _A\left(z\right)`$. In fact, $`\rho _{h,A}\left(z\right)`$ may take in general positive and negative values. It can be regarded as a generalized probability distribution function as long as the association between operators and functions is based on antinormal ordering,
$$Tr\left\{\widehat{\rho }G_A(A,A_h^{})\right\}=\rho _{h,A}\left(z\right)G_A(z,z^{})d^2z$$
(34)
In particular, for a diagonal density matrix $`\rho _{h,A}\left(z\right)=\rho _A\left(z\right)`$ while for $`\widehat{\rho }=|w,hw,h|`$
$`\rho _{h,A}\left(z\right)={\displaystyle \frac{1}{\pi }}{\displaystyle \frac{E_h\left(w^{}z\right)}{E_h\left(w^{}w\right)}}\mathrm{exp}\left[z^{}\left(wz\right)\right]`$
Consequently, the transformation (31) applies if there exists the Fourier transform of $`\mathrm{exp}\left[z^{}\left(wz\right)\right]E_h\left(w^{}z\right)`$. Analogously, working with $`\widehat{\rho }=|w,h^1w,h^1|`$ we arrive at the same conclusion for $`\mathrm{exp}\left[z^{}\left(wz\right)\right]E_{h^1}\left(w^{}z\right)`$. This in turn implies that $`E_{h^1}\left(w^{}z\right)`$ and $`E_h\left(w^{}z\right)`$ cannot grow at infinity as quickly as $`\mathrm{exp}\left(zz^{}\right)`$.
We recall that in the coherent states representation of a bounded operator $`\widehat{O}`$, the vanishing of $`z\left|\widehat{O}\right|z=0`$ in a domain of the complex plane of finite area implies the vanishing of $`\widehat{O}`$ itself (see Refs. ). Since $`d_h`$ has been assumed non singular the same theorem holds true for $`z\left|\widehat{O}_h\right|z`$, so that two deformed density matrices having the same function $`\rho _{h,A}\left(z\right)`$ over some area of $`z`$, must coincide. In conclusion, Eq. (33) establishes a one-to-one correspondence between the operator $`\widehat{\rho }`$ and the function $`\rho _{h,A}\left(z\right)`$.
When the deformation admits the unity resolution (24) a density matrix can be represented in several cases by a P-representation,
$$\widehat{\rho }=P_h\left(\alpha \right)\alpha ,h\alpha ,h๐\mu $$
(35)
in which $`P_h\left(\alpha \right)`$ can be regarded as a generalized probability distribution function as long as the association between operators and functions is based on normal ordering,
$$Tr\left\{\widehat{\rho }G_N(A^{},A)\right\}=P_h\left(\alpha \right)G_N(\alpha ,\alpha ^{})๐\mu $$
(36)
When $`\widehat{\rho }`$ is represented in the form (35) the master equation of $`\widehat{\rho }`$ can be in many cases transformed in a master equation for $`P_h.`$ This circumstance becomes particularly valuable in the study of the decay of an excited trapped ion toward the fundamental dark state. In this case we are faced for example with operators of the form
$`A^{}\widehat{\rho }`$ $`=`$ $`{\displaystyle P\left(\alpha \right)m_h\left(\alpha \alpha ^{}\right)h^2\left(\alpha \frac{}{\alpha }\right)\frac{}{\alpha }\alpha \alpha d^2\alpha }`$
$`=`$ $`{\displaystyle \alpha \alpha \frac{}{\alpha }h^2\left(1\alpha \frac{}{\alpha }\right)\left\{P\left(\alpha \right)m_h\left(\alpha \alpha ^{}\right)\right\}d^2\alpha }`$
use having been made of Eq. $`\left(\text{15}\right).`$ Expanding $`P\left(\alpha \right)m_h\left(\alpha \alpha ^{}\right)`$ in power series of $`\alpha ^m\left(\alpha ^{}\right)^n`$ we see that
$`{\displaystyle \frac{}{\alpha }}h^2\left(1\alpha {\displaystyle \frac{}{\alpha }}\right)\alpha ^m\left(\alpha ^{}\right)^n=mh^2\left(1m\right)\alpha ^{m1}\left(\alpha ^{}\right)^n`$
## VI Nonlinear coherent states on a circle
The above definition of NCS states (we will call them of order 1) can be extended to the eigenstates of the operators $`A_{N+1}`$ of a more general form $`A_{N+1}=a^{N+1}h\left(\widehat{n}\right)`$ and so the equation
$`A_{N+1}|\alpha ,h,N+1,q=\alpha |\alpha ,h,N+1,q`$
with $`N>0`$ is considered.
The eigenstate belonging to the eigenvalue $`\alpha `$ is $`N+1`$โfold degenerate and $`q`$ is an integer ranging from 0 to $`N`$. In terms of Fock states we have
$$\alpha |\alpha ,h,N+1,q=N_{\alpha ,h,q}\underset{l=0}{\overset{\mathrm{}}{}}\frac{\alpha ^{l\left(N+1\right)+q}}{\sqrt{\left(l\left(N+1\right)+q\right)!}\left[h\left(l\left(N+1\right)+q\right)\right]!}|l\left(N+1\right)+q$$
(37)
with the normalization factor
$`\left|N_{\alpha ,h,q}\right|^2={\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\left|\alpha \right|^{2l\left(N+1\right)+2q}}{\left(l\left(N+1\right)+q\right)!\left(\left[h\left(l\left(N+1\right)+q\right)\right]!\right)^2}}`$
where
$`\left[h\left(l\left(N+1\right)+q\right)\right]!=h\left(q\right)h\left(N+1+q\right)\mathrm{}h\left(l\left(N+1\right)+q\right).`$
Such a state can also be expressed as a sum of NCS (see Eq.13). In facts, by introducing the function $`h^{\left(N+1\right)}\left(n\right)`$ defined recursively by
$`h^{\left(N+1\right)}\left(l\left(N+1\right)+q\right)={\displaystyle \frac{h\left(q1\right)\left[h\left(l\left(N+1\right)+q\right)\right]!}{h\left(q\right)\left[h\left(l\left(N+1\right)+q1\right)\right]!}}h^{\left(N+1\right)}\left(q\right)`$
we have also
$$|\alpha ,h,N+1,q=N_{\alpha ,h,q}^{}\underset{k=0}{\overset{N}{}}\left(ฯต^{}\right)^{qk}|\alpha ฯต^k,h^{\left(N+1\right)}$$
(38)
with $`ฯต=\mathrm{exp}\left(\frac{i2\pi }{N+1}\right)`$ and $`N_{\alpha ,h,q}^{}`$ a normalization coefficient.
We have obtained that a NCS coherent state of order $`N+1`$ is decomposed in the sum of $`N+1`$ first order NCS of complex amplitudes $`\alpha ,`$ $`\alpha ฯต^q,\mathrm{},\alpha \left(ฯต^q\right)^N`$ distributed uniformly on a circle. These states, referred to as โcrystallized catsโ in Ref. , were introduced for the linear oscillator in the attempt to generalize the optical Schrรถdinger cats of harmonic oscillators.
Using the deformed displacement operator we have also
$`|\alpha ,h,N+1,q=N_{\alpha ,h,q}^{}\left({\displaystyle \underset{k=0}{\overset{N}{}}}\left(ฯต^{}\right)^{qk}๐_{h^{\left(N+1\right)}}\left(\alpha ฯต^k\right)\right)|0`$
In conclusion, the Hilbert space is the direct sum of $`N+1`$ spaces $`=_0_1\mathrm{}_N`$, $`\left(q=0,1,\mathrm{}N\right)`$, each of them having for basis the Fock states $`|\mathrm{}\left(N+1\right)+q`$, as $`l(0,\mathrm{}\mathrm{}).`$
For $`N=1`$ the fundamental states of $`_0`$ and $`_1`$ are respectively the even and odd Schrรถdinger cats. It will be shown in a following paper that when the radiative damping is negligible, the initial density matrix separate in the product of two matrices evolving respectively toward the even and odd Schrรถdinger cats.
### A Wigner function
The Wigner function relative to these states on a circle can be shown to be given for a generic integer $`N`$ by
$`W_{N+1}(\stackrel{~}{q},\stackrel{~}{p})=N_{\alpha ,h_{N+1},q}^2e^{\left|q+ip\right|^2}{\displaystyle \underset{ll^{}}{}}\left(\sqrt{2}\left(\stackrel{~}{q}i\stackrel{~}{p}\right)\right)^{\left(l^{}l\right)\left(N+1\right)}`$ (39)
$`\times `$ $`{\displaystyle \frac{\left(\alpha \right)^{l\left(N+1\right)+q}}{\left[h_{N+1}\left(l\left(N+1\right)+q\right)\right]!}}{\displaystyle \frac{\left(\alpha ^{}\right)^{l^{}\left(N+1\right)+q}}{\left[h_{N+1}\left(l^{}\left(N+1\right)+q\right)\right]!}}{\displaystyle \frac{L_{l\left(N+1\right)+q}^{\left(l^{}l\right)\left(N+1\right)}\left(2\left(\stackrel{~}{q}^2+\stackrel{~}{p}^2\right)\right)}{\left[l^{}\left(N+1\right)+q\right]!}}`$ (40)
Analogously for the HusimiโKano Q-function $`Q_{N+1,q}(\stackrel{~}{q},\stackrel{~}{p})=\left|\frac{\stackrel{~}{q}+i\stackrel{~}{p}}{\sqrt{2}}|\alpha ,q,h_{N+1}\right|^2`$
$$Q_{N+1,q}(\stackrel{~}{q},\stackrel{~}{p})=N_{\alpha ,h_{N+1},q}^2e^{\frac{\stackrel{~}{q}^2+\stackrel{~}{p}^2}{2}}\left|\underset{l=0}{\overset{\mathrm{}}{}}\frac{\left(\alpha \frac{\stackrel{~}{q}i\stackrel{~}{p}}{\sqrt{2}}\right)^{l\left(N+1\right)+q}}{\left[l\left(N+1\right)+q\right]!\left[h_{N+1}\left(l\left(N+1\right)+q\right)\right]!}\right|^2.$$
(41)
with $`|\frac{\stackrel{~}{q}+i\stackrel{~}{p}}{\sqrt{2}}`$ a coherent-state vector.
For $`N=1`$ these states reduce to even $`\left(q=0\right)`$ and odd $`\left(q=1\right)`$ Schrรถdinger cats . In Ref. is examined the squeezing and antibunching effects by using the function $`h_1(n)`$ introduced in for the NCS. We will see in the following (see Eq. (50)) that the nonlinear cats representing the dark state of a trapped ion are properly described by the deformation $`h_2\left(n\right)=L_{n2}^2\left(\eta ^2\right)/[n\left(n1\right)L_{n2}\left(\eta ^2\right)]`$.
In Fig. 1 we show the Wigner functions for nonlinear even Schrรถdinger cats of amplitude $`\alpha =3.5`$ (real) and different parameters $`\eta `$. In the linear case ($`\eta =0`$ Fig. 1-a) the quantum interference is localized around the origin. The two coherent gaussian peaks are circularly shaped.
For increasing $`\eta `$ the nonlinearity flattens the interference pattern while the central interference fringes, particularly their negative part, become more evident. This is essentially due to a reshaping of the coherent contribution peak from a gaussian-like nearly circular shape to an elliptical one with minor axis parallel to the direction connecting the two coherent peaks. These come closer to the origin and the region where the Wigner function is non-zero shrinks notably.
A further increase in $`\eta `$ causes the progressive coming closer and closer of the main peaks, while the interference fringes become more localized. For higher $`\eta `$ there are some interference fringes spreading over the two coherent peaks. This phenomenon is dominant for very high $`\eta `$ values (Fig. 1 b,c,d $`\eta =0.5`$) where the main peaks come into the interference region and the coherent character of the two states forming the cat is no more distinguishable. The interference area become larger than in the linear case and a circular symmetry of the interference pattern become evident.
In Fig. 2 we present two NCS on a circle formed by the superposition of 3 and 4 NCS ($`\alpha =3.5`$ and $`\eta =0.33`$).
## VII Dark states
Cirac et al first proposed in 1993 a scheme for preparing coherent squeezed states of motion in an ion trap based on the multichromatic excitation of a trapped ion. Using two waves with beat frequency equal to twice the trap frequency a โdark resonanceโ appears in the fluorescence emitted by the ion, the ion is placed in a squeezed state. Similar โdark statesโ produced by a bichromatic field with beat frequency equal to the trap frequency were studied by Vogel et al. in 1996 and identified as nonlinear coherent states.
We will consider in the following a beat frequency which is a generic multiple of the trap frequency $`\nu `$, and the ion dark state is described by a generalized coherent state on a circle. We will consider a bichromatic field of the type
$$A_{b,N+1}=f_{N+1}\left(\widehat{n}\right)a^{N+1}\alpha _{N+1}f_0\left(\widehat{n}\right).$$
(42)
for which the dark state satisfies the equation
$`\left(a^{N+1}{\displaystyle \frac{f_{N+1}\left(\widehat{n}N1\right)}{f_0\left(\widehat{n}N1\right)}}\alpha _{N+1}\right)|\psi _{dark}=0.`$
that is (see Eq. (37))
$`|\psi _{dark}=|\alpha _{N+1},h_{N+1},N+1,q`$
with
$$h_{N+1}(\widehat{n};\eta ^2)=\left(N+1\right)!\frac{L_{\widehat{n}N1}^{N+1}}{\left(\widehat{n}N\right)_{N+1}L_{\widehat{n}N1}}$$
(43)
and
$`\alpha _{N+1}={\displaystyle \frac{\mathrm{\Omega }_0}{\mathrm{\Omega }_{N+1}}}{\displaystyle \frac{\left(N+1\right)!}{\left(i\eta \right)^{N+1}}}`$
In short the dark state is the superposition of $`N+1`$ nonlinear coherent states which are equidistantly separated from each other along a circle with modulation factor $`ฯต^k=\mathrm{exp}\left(\frac{2\pi ik}{N+1}\right)`$.
In particular
$`h_1(\widehat{n};\eta ^2)`$ $`=`$ $`{\displaystyle \frac{L_{\widehat{n}1}^1}{\widehat{n}L_{\widehat{n}1}}}={\displaystyle \frac{L_{\widehat{n}1}L_{\widehat{n}}}{\eta ^2L_{\widehat{n}1}}}`$ (44)
$`h_2(\widehat{n};\eta ^2)`$ $`=`$ $`{\displaystyle \frac{2L_{\widehat{n}2}^2}{\widehat{n}\left(\widehat{n}1\right)L_{\widehat{n}2}}}`$ (45)
and
$`\alpha _1=i{\displaystyle \frac{\mathrm{\Omega }_0}{\mathrm{\Omega }_1}},\alpha _2={\displaystyle \frac{\mathrm{\Omega }_0}{\mathrm{\Omega }_2}}{\displaystyle \frac{2}{\eta ^2}}`$
Expressing the Laguerre polynomials by their asymptotic expression $`\sqrt{n\eta ^2}h_1(n;\eta ^2)`$ tends, for $`n\mathrm{}`$, to a function depending on the product $`n\eta ^2`$ only,
$$\sqrt{n\eta ^2}h_1(n;\eta ^2)=\mathrm{tan}\left(2\sqrt{n\eta ^2}\frac{\pi }{4}\right)+O\left(n^{3/4}\right)$$
(46)
Note the oscillating behavior of the eigenvalues $`E\left(n\right)\mathrm{tan}^2\left(2\sqrt{n\eta ^2}\frac{\pi }{4}\right)/\eta ^2`$. This circumstance implies that each eigenstate is encompassed by an infinite countable set of eigenstates of slightly different energies. For exceptional values of $`\eta `$ some eigenvalues can vanish. In these cases the series representation of the relative NCS looses its meaning. The behavior of $`E\left(n\right)`$ as $`n\mathrm{}`$ has strong implication on the resolution of unity, as we will see in the following.
As a consequence of (46), the logarithm of the factorial $`\left[h_1^2(n,\eta ^2)\right]!n!`$ times $`\eta ^{2n}`$ tends asymptotically to
$`\mathrm{log}(\left[h_1^2(n;\eta ^2)\right]!n!\eta ^{2n})\stackrel{n\mathrm{}}{}{\displaystyle \frac{\pi }{8\eta ^2}}{\displaystyle ^{\left(\frac{8}{\pi }\right)^2n\eta ^2}}\mathrm{log}\left[\mathrm{tan}\left({\displaystyle \frac{\pi }{4}}(\sqrt{z}1)\right)\right]^2dz+const`$
$`\stackrel{n\mathrm{}}{}{\displaystyle \frac{1}{\eta ^2}}u\left(\sqrt{n\eta ^2}\right)+const`$
where $`u\left(x\right)`$ is an oscillating entire function
$`u\left(x\right)`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left(4x\mathrm{}\left[Li_2\left(e^{i\phi }\right)+Li_2\left(e^{i\phi }\right)\right]+\mathrm{}\left[Li_3\left(e^{i\phi }\right)Li_3\left(e^{i\phi }\right)\right]\right)`$ (47)
$`=`$ $`{\displaystyle \frac{4}{\pi }}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}\left(1\right)^k{\displaystyle \frac{4x\left(2k+1\right)1}{\left(2k+1\right)^3}}\mathrm{cos}\left[\left(2k+1\right)4x\right]`$ (48)
with $`Li_n\left(z\right)=_{k=1}^{\mathrm{}}z^k/k^n`$ the polylogarithm function. According to (48) $`\mathrm{log}\left(\left[h_1^2(n,\eta ^2)\right]!n!\eta ^{2n}\right)`$ is an oscillating function of $`\sqrt{n\eta ^2}`$, with the envelope expanding proportionally to $`\sqrt{n\eta ^2}`$, as shown in Fig. 3. This behavior is confirmed by the exact expressions of $`\mathrm{log}\left(\left[h_1^2(n,\eta ^2)\right]!n!\eta ^{2n}\right)`$ plotted in Fig. 4 versus $`n`$ for $`\eta ^2=0.01,0.02,0.1`$ and $`0.2`$.
In particular there exists a countable infinite sequence of values of $`n\eta ^2`$ for which $`\left[h_1^2(n,\eta ^2)\right]!n!\eta ^{2n}`$ is close to $`\mathrm{exp}\left(\frac{16\sqrt{n}}{\pi \eta }\right)`$, so that the NCS can be normalized only for $`\left|\alpha \eta \right|`$ less than one. In other words, contrarily to the linear coherent states the NCS relative to a trapped ion fill the open circle $`1/\eta `$ in the complex plane. As $`\eta 0`$ the domain of existence tends to the whole complex plane, as for the linear coherent states. While these states are normalized for $`\alpha \eta `$ inside the unit circle, the scalar product of $`|\alpha _1,h`$ and $`|\alpha _2,h`$ is defined even if one the numbers $`\alpha _1`$or $`\alpha _2`$ has an arbitrary modulus as long as the product of the moduli is less than $`1/\eta ^2`$. A similar situation occurs for the harmonious states described by the deformation function $`h\left(n\right)=1/\sqrt{n}`$.
For generic combinations of vibrational excitation levels and parameter $`\eta ^2`$ the ion rovibronic dynamics fully displays its nonlinear character. An example of this feature has been seen above in connection with the discussion of the Wigner functions of some nonlinear Schrรถdinger cats and states on a circle of order 3 and 4.
### A Unity resolution
Being these NCS restricted to values of $`\alpha `$ such that $`\left|\alpha \eta \right|<1`$ the Mellin transform (26) reduces to the single-sided Laplace transform
$`g\left(s\right)=\underset{\epsilon 0}{lim}{\displaystyle _\epsilon ^{\mathrm{}}}m_h\left(\eta ^2e^t\right)e^{st}๐t`$
Then, $`g\left(s\right)`$ is the right-sided Laplace transform of the distribution $`m_h\left(\eta ^2e^t\right)`$ and tends asymptotically to the analytic function $`\mathrm{exp}\left[\eta ^2v\left(\sqrt{s\eta ^2}\right)\right]`$ of $`s`$ in the half-plane $`Re`$ $`s>1`$ and is bounded according to
$`g\left(s\right)Ke^{TRes}`$
with $`T`$ a negative infinitesimally small constant and $`K`$ a constant. Consequently $`m_h`$ is a distribution with support bounded on the left at $`t=T<0`$ (see Ref. , corollary 8.4-1a.). This means that it is necessary to include in the unity resolution un-normalizable states of amplitude $`\left|\alpha \right|>\eta ^1`$.
### B Approximate deformations
The difficulties in dealing with this deformation can be overcome by using approximate deformations. This is justified by the circumstance that in laser cooling experiments one deals with ions occupying a finite number of vibrational levels.
In particular when the parameter $`\eta ^2`$ is not very large the deformations $`h_{1,2}\left(n\right)`$ can be approximated by a few terms of the series expansion
$`h_1(\widehat{n};\eta ^2)`$ $`=`$ $`1+{\displaystyle \frac{\widehat{n}1}{2}}\eta ^2+{\displaystyle \frac{2\widehat{n}^23\widehat{n}+1}{6}}\eta ^4+{\displaystyle \frac{11\widehat{n}^322\widehat{n}^2+13\widehat{n}2}{48}}\eta ^6+\mathrm{}`$ (49)
$`h_2(\widehat{n};\eta ^2)`$ $`=`$ $`1+{\displaystyle \frac{2}{3}}\left(\widehat{n}2\right)\eta ^2+{\displaystyle \frac{11\widehat{n}^239\widehat{n}+34}{24}}\eta ^4+{\displaystyle \frac{19\widehat{n}^396\widehat{n}^2+159\widehat{n}86}{60}}\eta ^6+\mathrm{}`$ (50)
Approximating the deformation $`h_1\left(n\right)`$ by $`1+\frac{\widehat{n}1}{2}\eta ^2`$ we have that $`m_{\frac{\eta ^2}{2}h_1}\left(x\right)`$ coincides with the Mellin antitransform of $`\mathrm{\Gamma }\left(s\right)\mathrm{\Gamma }^2\left(\frac{2}{\eta ^2}+s1\right)`$. Before using Eq. (29) with $`A=3`$, $`B=0`$ and $`a_1=0`$, $`a_2=a_3=\frac{2}{\eta ^2}1`$, we have to remove the degeneracy $`a_2=a_3`$ by evaluating Eq.(29) for $`a_2a_3=\epsilon `$ and letting $`\epsilon 0`$. Since $`lim_{\epsilon 0}\mathrm{\Gamma }\left(\epsilon \right)+\mathrm{\Gamma }\left(\epsilon \right)2\mathrm{sinh}\left(\gamma \epsilon \right)/\epsilon =2\gamma `$ with $`\gamma =0.57721`$ the Eulerโs constant, then
$`a_2^2m_{h_1}\left(x\right)=\mathrm{\Gamma }^2\left[a_2\right]_0F_2[;1a_{2,}1a_2;{\displaystyle \frac{x}{a_2^2}}]2\gamma \mathrm{\Gamma }[a_2]\mathrm{\Gamma }\left[2a_2\right]_0F_2[;1,1+a;{\displaystyle \frac{x}{a_2^2}}]\left({\displaystyle \frac{x}{a_2^2}}\right)^{a_2}`$
At the same time the h-exponential reads
$`E_h\left(v\right)=_0F_3[;1,a_2+1,a_2+1;{\displaystyle \frac{v}{a_2^2}}]`$
Expanding the generalized factorial $`\left[h_1(\widehat{n},\eta ^2)\right]!`$ to the second order in $`\eta ^2`$ the NCS $`|\alpha ,h_1`$ can be expressed as a combination of excited coherent states
$`|\alpha ,h_1=\left(1+\eta ^2{\displaystyle \frac{\eta ^4}{12}}\right)|\alpha +\left({\displaystyle \frac{\eta ^2}{2}}+{\displaystyle \frac{7\eta ^4}{72}}\right)\alpha |\alpha ,1+\left({\displaystyle \frac{\eta ^2}{2}}+{\displaystyle \frac{\eta ^4}{24}}\right)\alpha ^2|\alpha ,2{\displaystyle \frac{5\eta ^4}{36}}\alpha ^3|\alpha ,3+O\left(\eta ^6\right)`$
A better approximation can be obtained by representing the Laguerre polynomials $`L_{n1}\left(\eta ^2\right)`$ by a finite sum of powers of $`\eta ^2,`$
$`L_n\left(\eta ^2\right){\displaystyle \underset{k=0}{\overset{K}{}}}\left(1\right)^k\left({\displaystyle \genfrac{}{}{0pt}{}{n}{k}}\right){\displaystyle \frac{\eta ^{2k}}{k!}}=P_K\left(n\right)`$
so that $`h_1(n;\eta ^2)`$ (see Eq. (45)) can be replaced by a rational function
$$h_{A,B}\left(n\right)=\frac{P_{A+1}\left(n1\right)P_{A+1}\left(n\right)}{\eta ^2P_B\left(n1\right)}=\gamma \frac{_{i=1}^A\left(n+a_i\right)}{_{j=1}^B\left(n+b_j\right)}$$
(51)
For using the characteristic function $`\rho _{h,A}`$ we should choose $`A=B`$, while for introducing the P-representation $`AB`$. In particular, Cirac et al and Blockley et al have expanded the exponential of Eq. (3) up to the second order in $`\eta `$, their case corresponds to $`A=B=4`$.
If the roots $`a_i`$ and $`b_i`$ are not integer we have (cf. Eq. (28))
$`\left[h_{AB}\left(s\right)\right]!=\gamma ^{s1}\mathrm{\Gamma }^1\left[{\displaystyle \genfrac{}{}{0pt}{}{\left(a\right)}{\left(b\right)}}\right]\mathrm{\Gamma }\left[{\displaystyle \genfrac{}{}{0pt}{}{\left(a\right)+s}{\left(b\right)+s}}\right]`$
For this class of deformations the measure is given by the combination of the generalized hypergeometric functions of Eq. (27), subject to the precaution of removing the degeneracy of the coefficients $`a_i,b_i`$.
In particular for $`A=B=1`$ the HusimiโKano Q-function (41) relative to the state $`|z,h`$ reduces to
$`Q_1(w)e^{\left|wz\right|^2/2}\left|\varpi ^\nu J_\nu \left(\varpi \right)\right|^2_{\left|\varpi \right|\mathrm{}}Ce^{\left|wz\right|^2/2}\left[1+{\displaystyle \frac{\mathrm{cot}\left(\nu \pi \right)}{2}}\left(\zeta w^{2\nu }+\zeta ^{}w^{2\nu }\right)\right]`$
with $`\nu =\frac{1}{\eta ^2}+\frac{1}{2}`$, $`\varpi =\frac{i}{2}zw^{}`$ and $`\zeta =\left(i\frac{ez}{4\nu }\right)^{2\nu }`$. Consequently, the projector $`|z,hz,h|`$ is represented in terms of the undeformed coherent states by a P-representation containing derivatives of the Dirac function of the very high order $`\nu `$. This confirms the advantage of using the representation (35).
For a finite rank density matrix $`m_h\left(x\right)`$ can be represented by a finite combination of Laguerre polynomials
$$m_h\left(x\right)=e^x\underset{n=0}{\overset{n_{\mathrm{max}}}{}}m_nL_n\left(x\right)$$
(52)
Imposing the condition (25) for $`0nn_{\mathrm{max}}`$ yields
$`m_n={\displaystyle \underset{m}{}}\left(1\right)^m\left({\displaystyle \genfrac{}{}{0pt}{}{n}{m}}\right)\left(\left[h\left(m\right)\right]!\right)^2`$
In Fig. 5 we have plotted these approximate measure functions for different values of $`\eta ^2(=0.015,0.0156,0.0158,0.016)`$, representing density operators relative to ions excited up to the level $`n=50`$. For these values of $`\eta ^2`$ the inclusion of a larger number of terms $`\left(n>50\right)`$ would lead to measures taking negative values.
## VIII Conclusions
The vibrational steady states of ions placed in a parabolic trap and driven by bichromatic fields detuned by multiples of the vibrational frequency provide a class of realizations of the nonlinear version of the so-called coherent states on a circle. The most well known example is that of the nonlinear Schrรถdinger cat states. As for the linear case also these states can be decomposed into finite sums of nonlinear coherent states, which can be considered as the building blocks of the vibrational wavefunctions of systems driven by laser fields detuned by multiples of the vibrational frequency.
This class of states is well described by the Wigner function, which has been computed for states on a circle of degree two (cats), three and four. The relative patterns show a dramatic dependence on the Lamb-Dicke parameter $`\eta `$, which measures the degree of departure from the linear case. This behavior is due to the irregular dependence of the ion deformation function on the Fock state index n.
With the aim of investigating the possibility of extending some mathematical tools of the linear coherent state theory to NCS it has been introduced a deformed displacement operator $`๐_h\left(\alpha \right)`$, which in analogy with the linear one generates the NCS $`|\alpha ,h`$ by displacing the Fock vacuum state $`\left(|\alpha ,h=๐_h\left(\alpha \right)|0\right)`$. The penalty paid for this extension is the loss of the unitarity. However, it allows the construction of a linear functional, which can be used for representing density operators by means of a generalized probability distribution function $`\rho _{A,h}\left(z\right)`$.
The peculiarities of NCS connected with trapped ions become evident when the deformation factorial is analyzed for very large $`n`$. The weight of the n-th Fock state contributing to a NCS exhibits an almost periodic behavior by taking very large and very small values of the order of $`e^{\pm C\sqrt{n\eta ^2}}`$. Consequently, the NCS can be normalized only for $`\alpha `$ filling the open circle $`1/\eta `$ in the complex $`\alpha `$-plane. In addition, this behavior prevents the existence of a regular measure for resolving the unity. These pathologies mark the difference with q-oscillators whose deformation function is an increasing function of $`n`$.
For extending to these NCS the P-representation formalism it is necessary to replace the deformation with an approximate one, for which there exists a measure. These approximate NCS can be constructed in different ways. Two examples are provided. In the first case the deformation is represented by a rational function of the occupation number, obtained by truncating the Laguerre polynomials to some order in $`\eta ^2`$. By increasing the degree of these rational functions it is possible to represent accurately the actual deformation for occupation numbers extending up to infinity. The respective measures are the Mellin antitransforms of Gamma function products $`\mathrm{\Gamma }\left[\genfrac{}{}{0pt}{}{\left(a\right)}{\left(b\right)}\right]`$ and are given by combinations of generalized hypergeometric functions. In the other case the measure is represented by a finite combination of Laguerre polynomials. In this way it is possible to represent exactly the factorials up to a given level $`n`$, although it is not possible to obtain in general a positive definite measure.
The NCS can provide a basis for studying the trapped ion evolution by representing the statistical expectation values of either antinormal $`G_A(A,A_h^{})`$ or normal $`G_N(A,A_h^{})`$products of $`A`$ and $`A_h^{}`$ as integrals of the probability distributions $`\rho _{h,A}\left(z\right)`$ or $`P_h\left(z\right)`$ times the classical functions $`G_A(z,z^{})`$ and $`G_N(z,z^{})`$. Another possibility consists in transforming the density matrix master equation in an equivalent equation for the P-representation. This problem will be addressed in a more systematic way in a forthcoming paper.
Before concluding, it is remarkable that the deformation used for the q-oscillators, the ancestors of the NCS, is based on the same transformation used by Heine a century ago for generalizing the Gauss hypergeometric function.
## Acknowledgments
V.I.M. thanks the University of Naples โFederico IIโ for the kind hospitality and the Russian Foundation for Basic Research for the partial support under Project No. 99-02-17753.
References
L. Biedenharn, J. Phys. A22, L873 (1989)
A. MacFarlane, J. Phys. A22, 4581 (1989); see also C. P. Sun and H.-C Fu, J. Phys. A22, L983 (1989); M. Chaichian and P. Kulish, Phys. Lett. B232, 72 (1990)
R. W. Gray and C. A. Nelson, J. Phys. A23, L945 (1990); A. J. Bracken, D. S. McAnally, R. B. Zhang and M. D. Gould, J. Phys. A24, 1379 (1991); B. Jurco, Lett. Math. Phys. 21, 51 (1991); C. A. Nelson, โNovel implications of the q-analogue coherent statesโ, in โSymmetries in Science VIโ, ed. B. Gruber, Plenum Press, N.Y., p.563, (1993).
V. I. Manโko, G. Marmo, E.C.G. Sudarshan, and F. Zaccaria, Physica Scripta 55, 528 (1997); V. I. Manโko G.Marmo, and F. Zaccaria in โSymmetries in Scienceโ, Ed. B. Gruber and M. Ramek, Plenum Press, N. Y. (1997);
J. ฤrnugelj, M. Martinis, and V. Mikuta-Martinis, Phys. Lett. A 188, 347 (1994); Phys. Lett. B 318, 227 (1993); Phys. Rev. A 50, 1785 (1994);
E. T. Jaynes and F. W. Cummings, Proc. IEEE 51, 89 (1963);
B. Buck and C. V. Sukumar, Phys. Lett. 81 A, 132 (1981); J. Phys. A 17, 885 (1984); V. Buzek, Phys. Rev. A 39, 3196 (1989), G. S. Agarwal, J. Opt. Soc. Am. B2, 480 (1985); C. C. Gerry, Phys. Rev. A 37, 2683 (1988); C. V. Sukumar and B. Buck, Phys. Lett. 83 A, 211 (1981); A. S. Shumovsky, Fam Le Kien and E. I. Aliskenderov, Phys. Lett. A 124, 351 (1987);
J. J. Cirac, R. Blatt, P. Zoller, and W. D. Phillips, Phys. Rev. A 46, 2668 (1992)
C. A. Blockley and D. F. Walls, Phys. Rev. A 47, 2115 (1993); C. A. Blockey, D. F. Walls, and H. Risken, Europhys. Lett. 17, 509 (1992)
J. H. Eberly, N. B. Narozhny, and J. J. Sanchez-Mondragon, Phys. Rev. Lett. 44, 1323 (1980); N. B. Narozhny, J. J. Sanchez, and J. H. Eberly, Phys. Rev. A 23, 236 (1981); P. Meystre and M. S. Zubairy, Phys. Lett. 89 A, 390 (1982); C. C. Gerry, Phys. Rev. A 37, 2683 (1988); J. R. Kukeinski and J. L. Madajczyk, Phys. Rev. A 37, 317 (1988);
F. Diedrich and H. Walter, Phys. Rev. Lett. 58, 203 (1987); M. Schubert, I. Siemers, R. Blatt, W. Neuhauser, and P. E. Toscheck, Phys. Rev. Lett. 68, 3016 (1992)
W. Vogel, J. Phys. B: At. Mol. Phys. 16, 4481 (1983); W. Vogel and Th. Ullmann, J. Opt. Soc. Am. B3, 441 (1986)
J. J. Cirac, R. Blatt, A. S. Parkins, and P. Zoller, Phys. Rev. Lett. 70, 556 (1993); J. I. Cirac, P. Zoller, Phys. Rev. Lett. 74, 4091 (1995).
R. L. de Matos Filho and W. Vogel, Phys. Rev. A49, 2812 (1994); Phys. Rev. A 54, 4560 (1996);
S. Mancini, Phys. Lett. A 233, 291 (1997).
X-G Wang and H-C Fu, quant-ph/9903013 vs. Nov (1999).
V. Bargmann, Commun. Pure and Appl. Math. 14, 187 (1961).
C. L. Mehta and E.C.G. Sudarshan, Phys. Rev. 138B, 274 (1965); see also J. R. Klauder, J. Math. Phys. 5, 177 (1964) and T. F. Jordan, Phys. Lett. 11, 289 (1964)
F. Diedrich, J. C. Bergquist, Wayne M. Itano, and D. J. Wineland, Phys. Rev. Lett. 62, 403 (1989).
G.S. Agarwal and K. Tara, Phys. Rev. A 43, 492 (1991); V. V. Dodonov, Ya. A. Korennoy, and V. I. Manโko and Y.A. Moukhin, Quantum and Semiclassical Optics 8, 413 (1996);
S. Sivakumar, quant-ph/9806061;
R. J. Glauber, Phys. Rev. Lett., 10, 84 (1965); E. C. G. Sudarshan, Phys. Rev. Lett. 10, 277 (1965);
E. C. G. Sudarshan, Int. J. Theor. Phys. 32, 1069 (1993)
\[24 \] L. J. Slater, โGeneralized Hypergeometric Functionsโ, Cambridge University Press, Cambridge (1966); Proc. Camb. Phil. Soc. 51, 577 (1955)
Jinzuo Sun, Jisuo Wang, and Chuankui Wang, Phys. Rev. A 44, 3369 (1991);
O. Castaรฑos, R. Lopรฉz-Peรฑa, V.I. Manโko, J.Russ. Laser Res., 16, 477 (1995)
J. Janszky, P. Domokos, and P. Adam, Phys. Rev. A 48, 2213 (1994); Ren-Shan Gong, Phys. Lett. A 233, 297 (1997).
E. Wigner, Phys. Rev., 40, 749 (1932); K. Husimi, Proc. Phys. Math. Soc (1965).
B. Roy, Phys. Lett. A 249, 25 (1998); B.Roy and P. Roy, Phys. Lett. A 263, 48 (1999)
A. H. Zemanian, โDistribution Theory and Transform Analysisโ, McGraw-Hill, New York (1965).
E. Heine, Handuch die Kugelfunctionen, (1898) quoted by Slater (Ref.).
Figure captions
Fig. 1: Top and side views of the Wigner function (Eq. 39) relative to linear (a) and nonlinear (b,c,d) even Schrรถdinger cat states for different values of $`\alpha =3.5`$ (real) and two different values of $`\eta `$. (a) $`\eta =0`$ (linear case), the two coherent peaks are almost circularly shaped the interference pattern is axially symmetric along the axis defined by the center of the coherent peaks; (b โ top-view, c โ side-view and d โ air-view) $`\eta =0.5`$, the gaussian peaks are no more distinguishable from the circularly symmetric interference pattern.
Fig. 2: Top view of the Wigner function (Eq. 39) for 3 and 4 NCS states sitting on the circle. $`\alpha =3.5`$ (real) and $`\eta =0.33`$. This value of the nonlinearity $`\eta `$ leads to a smooth reshaping of the coherent peaks from a nearly circular shape to an elliptic one.
Fig. 3 Asymptotic expression of $`u\left(\sqrt{z}\right)`$ versus $`z`$ (see Eq. (48))
Fig. 4 $`\mathrm{log}\left(\left[h_1^2(n,\eta ^2)\right]!n!\eta ^{2n}\right)`$ versus $`n`$ for $`\eta ^2=0.01`$ (a),$`0.02`$ (b)$`,0.1`$ (c) and $`0.2`$ (d).
Fig. 5 Measure $`m_h\left(x\right)`$ versus x obtained as a combination of 50 Laguerre polynomials (Eq. (52)) and four different values of $`\eta ^2=0.015,0.0156,0.0158,0`$.$`016`$ |
warning/0003/hep-th0003267.html | ar5iv | text | # Dimensional Transmutation and Dimensional Regularization in Quantum Mechanics II. Rotational Invariance
## I INTRODUCTION
Dimensional transmutation has become a standard concept for the analysis of quantum field theories devoid of intrinsic dimensional parameters. In the first paper in this series , we have shown that dimensional transmutation is a more general phenomenon related to dimensional analysis and renormalization, and we provided a general theory for its description in nonrelativistic quantum mechanics. The main purpose of this paper is to apply the general theory developed in Ref. to an exhaustive analysis of examples that exhibit rotational invariance. Parenthetically, as shown in Ref. , scale-invariant potentialsโthose that possess scale symmetry at the classical level and do not exhibit any explicit dimensional scaleโare exactly described by homogeneous functions of degree $`2`$; this class includes contact potentials as well as ordinary potentials of the form $`1/r^2`$, with a possible angular dependence. Thus, within the subset of such potentials with rotational invariance, the following two problems stand out: the two-dimensional delta-function potential and the inverse square potential . In addition to solving these potentials, we will expand the general framework presented in in order to encompass further understanding of the scattering sectorโincluding a partial-wave analysisโas well as a reexamination of the peculiar boundary conditions satisfied by these potentials at the origin.
The plan of this paper is as follows. In Section II we discuss the two-dimensional delta-function potential and compare the results with those earlier derived in Ref. . In Section III we summarize the known properties of the inverse square potential (but generalizing them to $`D`$ dimensions) and examine the issue of the boundary condition at the origin; the conclusion of that section is that renormalization is needed in the strong-coupling regime. In Section IV we provide the required dimensional renormalization of the inverse square potential and find a complete solution to the problem by means of a duality transformation. Section V summarizes our main conclusions and outlines a strategy for answering additional questions. The appendices explicitly use the concept of rotational invariance and deal with the $`D`$-dimensional central-force problem, the $`D`$-dimensional partial-wave expansion, and the duality transformation.
We conclude this introduction with some comments about notation and background, with which we assume that the reader has some familiarity from Ref. . In particular, in this paper, we will solve the Schrรถdinger equation associated with
$$V(๐ซ)=\lambda W(๐ซ),$$
(1)
for the two-dimensional delta-function potential
$$W(๐ซ)=\delta ^{(2)}(๐ซ)$$
(2)
(Section II), and for the inverse square potential
$$W(๐ซ)=\frac{1}{r^2}$$
(3)
(Sections III and IV). The regularized solutions will be obtained with dimensional continuation from $`D_0`$ to $`D`$ dimensions and will depend on the parameter $`ฯต=D_0D`$; from these solutions, we will identify the energy generating function $`\mathrm{\Xi }(ฯต)`$, by comparison with the master eigenvalue equation
$$\lambda \mu ^ฯต\mathrm{\Xi }(ฯต)|E(ฯต)|^{ฯต/2}=1,$$
(4)
where $`\lambda `$ is the dimensionless coupling constant, $`E`$ the energy, and $`\mu `$ the renormalization scale that arises from the bare coupling $`\lambda _B=\mu ^ฯต\lambda `$. The ensuing analysis of existence of bound states and related concepts will often refer to properties of the energy generating function $`\mathrm{\Xi }(ฯต)`$ and of the critical couplings
$$\lambda _n^{()}=\left[\underset{ฯต0}{lim}\mathrm{\Xi }_n(ฯต)\right]^1,$$
(5)
as discussed in Ref. , even though the derived formulas can mostly be analyzed on their own right and, for the most part, this paper is essentially self-contained.
## II TWO-DIMENSIONAL DELTA-FUNCTION POTENTIAL
The delta-function potential belongs to the class of pseudopotentials that arises in the low-energy limit of effective quantum field theory . As in Ref. , in this section we will use dimensional regularization and will focus only on those features that are associated with the dimensional transmutation of the two-dimensional representative of this class, Eqs. (1) and (2).
The two-dimensional delta-function potential has been extensively studied in the literature using a variety of regularization techniques: (i) momentum-cutoff regularization , (ii) real-space short-distance regularization (using a circular-well potential) , (iii) Pauli-Villars regularization , (iv) dimensional regularization , and (v) method of self-adjoint extensions . As expected, the final renormalized results are independent of the regularization scheme, despite the dissimilar appearance of the regularized expressions. Even though our treatment will be solely based on dimensional regularization, the establishment of a peculiar boundary condition at the origin shows a natural connection with method (v), where the singular nature of the potential at the origin is replaced by a boundary condition in order to preserve the self-adjoint character of the Hamiltonian.
The dimensionally regularized Schrรถdinger equation for the potential (2) can be solved in a number of different ways. In Ref. we solved it in momentum spaceโa procedure that works due to the zero-range nature of this potential . The same feature permits an effective solution in hyperspherical coordinates as if it were a central potential (according to the framework developed in Appendix A). In fact, the similarities between the two-dimensional delta-function and inverse square potentials will become more transparent, by the use of hyperspherical coordinates for both.
According to the dimensional-continuation prescription of Ref. , we Fourier transform Eq. (2) in $`D_0=2`$ dimensions, then perform a dimensional jump to $`D`$ dimensions in Fourier space, and finally return to position space in $`D`$ dimensions, with the obvious result
$$[_{๐ซ,D}^2\lambda \mu ^ฯต\delta ^{(D)}(๐ซ)]\mathrm{\Psi }(๐ซ)=E\mathrm{\Psi }(๐ซ).$$
(6)
From the generalized angular-momentum analysis for central potentials of Appendix A, it follows that the radial part of Eq. (6) in hyperspherical coordinates reduces to the homogeneous form
$$\left[\frac{d^2}{dr^2}+E\frac{(l+\nu )^21/4}{r^2}\right]u_l(r)=0,$$
(7)
for any $`r0`$. In Eq. (7), as well as in our subsequent analysis, both for the two-dimensional delta-function and inverse square potentials, the number $`D`$ of dimensions will usually appear in terms of the variable
$$\nu =D/21,$$
(8)
which will thereby simplify the form of most formulas.
Equation (7) is indistinguishable from a free particle except for the stringent boundary condition enforced by the delta-function singularity at the origin. We now move on to analyze this boundary condition.
### A Boundary Condition at the Origin for the Two-Dimensional Delta-Function Potential
Simple inspection of Eqs. (6) and (7) shows that the delta-function singularity represents the only difference between them. Even though the correct equation is (6), one can still safely use (7), provided that the singularity be replaced by an appropriate boundary condition at the origin. This can be established by means of the small-argument behavior of the wave function (neglecting the energy term as $`r0`$)
$$_{๐ซ,D}^2\mathrm{\Psi }(๐ซ)\stackrel{\left(r0\right)}{}\lambda \mu ^ฯต\mathrm{\Psi }(\mathrm{๐})\delta ^{(D)}(๐ซ),$$
(9)
which can be evaluated by comparison with the identity
$$_{๐ซ,D}^2\left[\frac{r^{(D2)}}{(D2)\mathrm{\Omega }_D}\right]=\delta ^{(D)}(๐ซ).$$
(10)
Equations (9) and (10) imply that the general form of $`\mathrm{\Psi }(๐ซ)`$ near the origin is
$$\mathrm{\Psi }(๐ซ)\stackrel{\left(r0\right)}{}\mathrm{\Psi }(\mathrm{๐})\frac{\lambda \mu ^ฯต}{(D2)\mathrm{\Omega }_D}r^{(D2)}+\mathrm{\Psi }_h(๐ซ),$$
(11)
where $`\mathrm{\Psi }_h(๐ซ)`$ is the general solution of the homogeneous equation. Moreover, Eq. (11) can be made more explicit by means of
$$\mathrm{\Psi }_h(๐ซ)=\underset{l=0}{\overset{\mathrm{}}{}}\left[C_l^{(+)}r^l+C_l^{()}r^{(l+2\nu )}\right]Y_L(\mathrm{\Omega }^{(D)})$$
(12)
and
$$_{๐ซ,D}^2\left[r^{(l+2\nu )}Y_L(\mathrm{\Omega }^{(D)})\right]^l\delta ^{(D)}(๐ซ),$$
(13)
with $`^l`$ being a certain linear combination of $`l`$th order derivatives; then, it follows that $`C_l^{()}=0`$ for all $`l`$, for otherwise $`C_l^{()}`$ would modify the inhomogeneous part of Eq. (11). Then, resolving the wave function into individual angular momentum components,
$$\mathrm{\Psi }(๐ซ)=\underset{l=0}{\overset{\mathrm{}}{}}R_l(r)Y_L(\mathrm{\Omega }^{(D)}),$$
(14)
we have the following set of boundary conditions. First, for $`l=0`$, the value of the constant $`C_0^{(+)}`$ corresponds to $`\mathrm{\Psi }(\mathrm{๐})`$, as required by self-consistency of the value of the wave function at the origin; then,
$$R_0(r)\stackrel{\left(r0\right)}{}R_0(0)\left[\frac{\lambda \mu ^{2\nu }\mathrm{\Gamma }(\nu )}{4\pi ^{\nu +1}}r^{2\nu }+1\right].$$
(15)
In addition, for $`l0`$,
$$R_l(r)\stackrel{\left(r0\right)}{}C_l^{(+)}r^l.$$
(16)
Equations (15) and (16) are the required boundary conditions at the origin. In particular, Eq. (15) implies that $`D<2`$ is the condition for regularity, while the case $`D=2`$ is โcritical.โ
### B Bound-State Sector for a Two-Dimensional Delta-Function Potential
The coupling $`\lambda `$ in Eq. (2) defines an โattractiveโ zero-range potential for $`\lambda >0`$ and a โrepulsiveโ one for $`\lambda <0`$. By the form of the potential, this physical argument implies that all states with $`E>0`$ are of the scattering type, while states with $`E<0`$ can only be bound and are impossible for $`\lambda <0`$. In other words, there exists a critical coupling
$$\lambda ^{()}=0,$$
(17)
which separates the theory into two regimes. Therefore, for a delta-function potential, the strong-coupling regime ($`\lambda >\lambda ^{()}=0`$) coincides with the set of attractive potentials, while the weak-coupling regime ($`\lambda <\lambda ^{()}=0`$) amounts to repulsive potentials.
Let us now consider the bound-state sector, with energy $`E=\kappa ^2<0`$, for an attractive two-dimensional delta-function potential. The corresponding solution of Eq. (7) is
$$\frac{u_l(r)}{\sqrt{r}}=\mathbf{\{}\text{ }I_{l+\nu }(\kappa r)\text{,}\text{ }K_{l+\nu }(\kappa r)\mathbf{\}}\text{ },$$
(18)
where the symbol $`\mathbf{\{}`$ , $`\mathbf{\}}`$ stands for linear combination, and $`I_p(z)`$ and $`K_p(z)`$ are the modified Bessel functions of the first and second kinds, respectively .
Of course, the boundary conditions restrict the selection in Eq. (18). First, the boundary condition at infinity leads to the rejection of the modified Bessel function of the first kind, so that (from Eq. (A15))
$$R_l(r)=A_lr^\nu K_{l+\nu }(\kappa r).$$
(19)
Next, the boundary condition at the origin, Eqs. (15) and (16), can be enforced at the level of Eq. (19) by considering the small-argument behavior of the modified Bessel function of the second kind ,
$$K_p(z)\stackrel{(z0)}{}\frac{1}{2}\left[\mathrm{\Gamma }(p)\left(\frac{z}{2}\right)^p+\mathrm{\Gamma }(p)\left(\frac{z}{2}\right)^p\right]\left[1+O(z^2)\right].$$
(20)
For $`l0`$, Eq. (19) gives a singular term proportional to $`r^p`$, with $`p=l+\nu `$, according to Eq. (20). Therefore, the boundary condition can only be satisfied for
$$l=0,$$
(21)
so that the delta-function potential, being of zero range, can only sustain bound states in the absence of a centrifugal barrier ($`s`$ states). Then, for the radial wave function with $`l=0`$,
$`R_0(r)`$ $`=`$ $`A_0r^\nu K_\nu (\kappa r)`$ (22)
$`\stackrel{\left(r0\right)}{}`$ $`{\displaystyle \frac{A_0}{2}}\left({\displaystyle \frac{\kappa }{2}}\right)^\nu \left[\mathrm{\Gamma }(\nu )+\mathrm{\Gamma }(\nu )\left({\displaystyle \frac{\kappa r}{2}}\right)^{2\nu }\right]\left[1+O(r^2)\right].`$ (23)
In conclusion, compatibility of Eqs. (15) and (23) requires that the following two conditions be simultaneously met: (i) that
$$A_0=2R_0(0)\left(\frac{2}{\kappa }\right)^\nu \frac{1}{\mathrm{\Gamma }(\nu )},$$
(24)
which just enforces the condition of finiteness at the origin and requires $`D<2`$ for regularity; and (ii) that the bound-state energies $`E`$ satisfy the eigenvalue condition
$$\frac{\lambda \mu ^{2\nu }}{4\pi }\left(\frac{|E|}{4\pi }\right)^\nu \mathrm{\Gamma }\left(\nu \right)=1.$$
(25)
Equations (23), (24), and (25) are in complete agreement with the results obtained in Ref. , to which we refer the reader for additional comments.
As there is no other quantum number in this problem, and $`l=0`$ is required, we see that Eq. (22) represents the ground state of the regularized system. Parenthetically, the proportionality constant $`A_0`$ can be found by normalization by means of the identity
$$_0^{\mathrm{}}x\left[K_\nu (x)\right]^2๐x=\frac{1}{2}\mathrm{\Gamma }(1+\nu )\mathrm{\Gamma }(1\nu ),$$
(26)
so that
$$\mathrm{\Psi }(๐ซ)=\frac{\kappa }{\pi ^{(\nu +1)/2}[\mathrm{\Gamma }(1\nu )]^{1/2}}\frac{K_\nu (\kappa r)}{r^\nu }.$$
(27)
Equations (25) and (27) indeed reduce to the known expressions for $`D=1`$ . Here, no attempt is made to draw any conclusions about the cases $`D>2`$, as they require a separate regularization procedure.
As discussed in Ref. , Eq. (25) is equivalent to the master eigenvalue equation (4), with an energy generating function
$$\mathrm{\Xi }(ฯต)=\frac{1}{4\pi }\left(4\pi \right)^{ฯต/2}\mathrm{\Gamma }\left(\frac{ฯต}{2}\right),$$
(28)
which, not having any discrete labels, produces just a ground state
$$E_{_{(\mathrm{gs})}}=\mu ^2e^{g^{(0)}}\mu ^2,$$
(29)
but no excited states. In Eq. (29), the symbol $``$ refers to the freedom to make the choice $`g^{(0)}=0`$, which means no loss of generality, as both $`g^{(0)}`$ and $`\mu `$ are totally arbitrary . Moreover, the critical coupling is $`\lambda ^{()}=0`$, as anticipated earlier by the argument leading to Eq. (17); more precisely, the coupling constant is asymptotically critical according to the scheme
$$\lambda (ฯต)=2\pi ฯต\left\{1+\frac{ฯต}{2}\left[g^{(0)}\left(\mathrm{ln}4\pi \gamma \right)\right]\right\},$$
(30)
where $`\gamma `$ is the Euler-Mascheroni constant.
Finally, the ground state wave function of the renormalized two-dimensional delta-function potential can be obtained in the limit $`\nu 0`$ of Eq. (27), and using Eq. (29), with the result
$$\mathrm{\Psi }_{_{(\mathrm{gs})}}(๐ซ)=\frac{\mu }{\sqrt{\pi }}K_0(\mu r),$$
(31)
which was derived by using a renormalized path integral in Ref. .
### C Scattering Sector for a Two-Dimensional Delta-Function Potential
Scattering under the action of a two-dimensional delta-function potential will also be analyzed by an explicit partial-wave resolution in hyperspherical coordinates according to the theory of Appendix B. The first step is to solve Eq. (7), which for the scattering sector, $`E=k^2>0`$, admits the solution
$$u_l(r)=\sqrt{r}\left[A_l^{(+)}H_{l+\nu }^{(1)}(kr)+A_l^{()}H_{l+\nu }^{(2)}(kr)\right],$$
(32)
where the $`H_p^{(1,2)}(z)`$ are Hankel functions, whose behavior near the origin implies the asymptotic dependence
$$u_l(r)\stackrel{\left(r0\right)}{}\frac{\sqrt{r}}{\pi i}\left\{\left[A_l^{(+)}e^{i\pi (l+\nu )}A_l^{()}e^{i\pi (l+\nu )}\right]\mathrm{\Gamma }(\nu )\left(\frac{kr}{2}\right)^{l+\nu }+\left[A_l^{(+)}A_l^{()}\right]\mathrm{\Gamma }(\nu )\left(\frac{kr}{2}\right)^{(l+\nu )}\right\}.$$
(33)
Equations (33) and (A15) yield
$`R_l(r)`$ $`\stackrel{\left(r0\right)}{}`$ $`{\displaystyle \frac{A^{()}}{\pi i}}\left[S_l^{(D)}(k)e^{i\pi (l+\nu )}e^{i\pi (l+\nu )}\right]\mathrm{\Gamma }(\nu )\left({\displaystyle \frac{kr}{2}}\right)^l`$ (34)
$`\times `$ $`\left\{1+{\displaystyle \frac{\left[S_l^{(D)}(k)1\right]\mathrm{\Gamma }(\nu )\left(2/k\right)^{2\nu }}{\left[S_l^{(D)}(k)e^{i\pi (l+\nu )}e^{i\pi (l+\nu )}\right]\mathrm{\Gamma }(\nu )}}r^{\left(2l+2\nu \right)}\right\},`$ (35)
where the scattering matrix elements $`S_l^{(D)}(k)`$ have been explicitly introduced by means of Eq. (B13).
Equation (35) should be compared again with the boundary condition at the origin. According to Eq. (13), for $`l0`$, the boundary condition (16) cannot be satisfied, unless
$$S_l^{(D)}(k)|_{l0}=1;$$
(36)
in particular, for the phase shifts,
$$\delta _l^{(D)}(k)|_{l0}=0,$$
(37)
with the conclusion that there is no scattering for $`l0`$. On the other hand, for the $`s`$ wave, Eqs. (15) and (35) yield a scattering matrix element
$$S_0^{(D)}(k)=\frac{1\lambda e^{i\pi \nu }\mathrm{\Xi }(2\nu )\left(E/\mu ^2\right)^\nu }{1\lambda e^{i\pi \nu }\mathrm{\Xi }(2\nu )\left(E/\mu ^2\right)^\nu },$$
(38)
where
$$\mathrm{\Xi }(2\nu )=\frac{\mathrm{\Gamma }(\nu )}{(4\pi )^{\nu +1}}$$
(39)
is identical to the energy generating function, Eq. (28). Equation (38) provides the phase shift $`\delta _0^{(D)}(k)`$ through Eq. (B7), whence
$$\mathrm{tan}\delta _0^{(D)}(k)=\mathrm{tan}\pi \nu \left[1\frac{\mathrm{sec}\pi \nu }{\lambda \left(k/\mu \right)^{2\nu }\mathrm{\Xi }(2\nu )}\right]^1.$$
(40)
We are now ready to derive the expressions for the two-dimensional delta-function potential. First, for $`\lambda <0`$ (repulsive potential), $`\lambda `$ remains unrenormalized and the limit $`ฯต=2\nu 0`$ yields
$$S_0^{(D)}(k)|_{\lambda <0}=1,$$
(41)
so that there is no scattering whatsoever for repulsive two-dimensional delta-function potentials. Instead, for $`\lambda >0`$ (attractive potential), in the limit $`ฯต=2\nu 0`$ these expressions should be renormalized by means of Eq. (30), whence
$$S_0^{(2)}(k)|_{\lambda >0}=\frac{\mathrm{ln}\left(E/\mu ^2\right)g^{(0)}+i\pi }{\mathrm{ln}\left(E/\mu ^2\right)g^{(0)}i\pi }.$$
(42)
Equation (42) can be further simplified with Eq. (29), leading to
$$S_0^{(2)}(k)=\frac{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)+i\pi }{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)i\pi },$$
(43)
and
$$\mathrm{tan}\delta _0^{(2)}(k)=\frac{\pi }{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)},$$
(44)
where, from this point on, it will be understood that scattering is nontrivial only for $`\lambda >0`$. In particular, from Eqs. (36), (43), (B5), (B8), and (B9), the scattering amplitude becomes
$`f_k^{(2)}(\mathrm{\Omega }^{(2)})`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{\pi k}}}{\displaystyle \frac{S_0^{(2)}(k)1}{2i}}`$ (45)
$`=`$ $`\sqrt{{\displaystyle \frac{2\pi }{k}}}\left[\mathrm{ln}\left({\displaystyle \frac{k^2}{E_{_{(\mathrm{gs})}}}}\right)i\pi \right]^1.`$ (46)
Equation (46) is identical to the corresponding expression derived in Ref. and coincides with the corresponding expressions derived in Refs. .
A number of consequences follow from Eqs. (43), (44), and (46):
1. The unique pole of the scattering matrix (43) corresponds to the unique bound state.
2. The phase shifts can be alternatively renormalized using a floating scale $`\mu `$, in the form
$$\frac{1}{\mathrm{tan}\delta _0^{(2)}(k)}=\frac{1}{\mathrm{tan}\delta _0^{(2)}(\mu )}+\frac{1}{\pi }\mathrm{ln}\left(\frac{k}{\mu }\right)^2.$$
(47)
3. Equation (46) yields a differential scattering cross section
$$\frac{d\sigma ^{(2)}(k,\mathrm{\Omega }^{(2)})}{d\mathrm{\Omega }_2}=\frac{2\pi }{k}\left\{\left[\mathrm{ln}\left(\frac{k^2}{E_{_{(\mathrm{gs})}}}\right)\right]^2+\pi ^2\right\}^1.$$
(48)
4. The total scattering cross section becomes
$$\sigma _2(k)=\frac{4\pi ^2}{k}\left\{\left[\mathrm{ln}\left(E/|E_{_{(\mathrm{gs})}}|\right)\right]^2+\pi ^2\right\}^1.$$
(49)
5. All the relevant quantities are logarithmic with respect to the energy and agree with the predictions of generalized dimensional analysis .
6. Equations (43), (44), (46), (48), and (49) relate the bound-state and scattering sectors of the theory and confirm that the two-dimensional delta-function potential is renormalizable.
## III INVERSE SQUARE POTENTIAL: INTRODUCTION
It is well known that the inverse square potential, $`V(๐ซ)r^2`$, displays a number of unusual features in its quantum-mechanical version. In a sense, it represents the boundary between regular and singular power-law potentials (Appendix C). As we will see in this section, its marginally singular nature can be traced back to its interplay with the centrifugal barrier and produces two regimes separated by a critical coupling. These features have become part of the standard background on central potentials in quantum mechanics . The difficulties encountered in the strong-coupling regime correspond to the classical picture of the โfall of the particle to the centerโ . It should be pointed out that this problem is relevant in polymer physics , as well as in molecular physics, where it appears in a modified form as a dipole potential .
In this section, our goal is to review the origin of this singular behavior and pave the way for the regularization and renormalization of the theory, which we will implement in Section IV.
### A Exact Solution for the Unregularized Inverse Square Potential
Stationary eigenstates of energy and orbital angular momentum of the unregularized inverse square potential are of the factorized form (A5), with an effective radial function (A15) that satisfies the $`D_0`$-dimensional radial Schrรถdinger equation (Eqs. (A12)โ(A18)),
$$\left[\frac{d^2}{dr^2}+E\frac{\left(l+\nu _0\right)^2\lambda 1/4}{r^2}\right]u_l(r)=0,$$
(50)
where $`\nu _0=D_0/21`$ and $`\lambda >0`$ corresponds to an attractive potential. Then solutions to Eq. (50) are of the form
$$u_l(r)=\sqrt{r}Z_{s_l}\left(\sqrt{E}r\right),$$
(51)
where $`Z_{s_l}(z)`$ represents an appropriate linear combination of Bessel functions of order
$$s_l=\sqrt{\lambda _l^{()}\lambda },$$
(52)
with
$$\lambda _l^{()}=\left(l+\nu _0\right)^2.$$
(53)
It is immediately apparent that the inverse square potential in Eq. (50) has the same dependence on the radial variable as the centrifugal potential, so that its effect is solely to modify the strength of the centrifugal barrier; correspondingly, the solution (51) is essentially a free-particle wave function where the replacement
$$(l+\nu _0)s_l$$
(54)
has been made. The parameter $`\lambda _l^{()}`$ in Eq. (53) plays the role of a critical coupling, i.e., the nature of the solutions changes abruptly around the value $`\lambda =\lambda _l^{()}`$, for any state with angular momentum $`l`$. Thus, $`\lambda _l^{()}`$ acts as the boundary between two regimes: (i) weak coupling, characterized by $`\lambda <\lambda _l^{()}`$, for which the order $`s_l`$ is real; and (ii) strong coupling, characterized by $`\lambda >\lambda _l^{()}`$, for which the order $`s_l=i\mathrm{\Theta }_l`$ is imaginary, with
$$\mathrm{\Theta }_l=\sqrt{\lambda \lambda _l^{()}}.$$
(55)
The character of the solutions also depends on the other relevant parameter in Eq. (50), namely, the energy $`E`$, in such a way that: (i) scattering states are only possible if $`E=k^2>0`$, with the argument of the Bessel functions in Eq. (51) being $`kr`$ (real); while (ii) bound states are only possible if $`E=\kappa ^2<0`$, with the argument of the Bessel functions in Eq. (51) being $`kr=i\kappa r`$ (imaginary). In conclusion, the solutions (51) fall into the following four families, according to the nature of the two relevant variables $`s_l`$ and $`E`$:
1. Bound-state sector of the weak-coupling regime, with
$$\frac{u_l(r)}{\sqrt{r}}=\mathbf{\{}\text{ }I_{s_l}(\kappa r)\text{,}\text{ }K_{s_l}(\kappa r)\mathbf{\}}\text{ };$$
(56)
2. Scattering sector of the weak-coupling regime, with
$$\frac{u_l(r)}{\sqrt{r}}=\mathbf{\{}\text{ }H_{s_l}^{(1)}(kr)\text{,}\text{ }H_{s_l}^{(2)}(kr)\mathbf{\}}\text{ };$$
(57)
3. Bound-state sector of the strong-coupling regime, with
$$\frac{u_l(r)}{\sqrt{r}}=\mathbf{\{}\text{ }I_{i\mathrm{\Theta }_l}(\kappa r)\text{,}\text{ }K_{i\mathrm{\Theta }_l}(\kappa r)\mathbf{\}}\text{ };$$
(58)
4. Scattering sector of the strong-coupling regime, with
$$\frac{u_l(r)}{\sqrt{r}}=\mathbf{\{}\text{ }H_{i\mathrm{\Theta }_l}^{(1)}(kr)\text{,}\text{ }H_{i\mathrm{\Theta }_l}^{(2)}(kr)\mathbf{\}}\text{ };$$
(59)
where the symbol $`\mathbf{\{}`$ , $`\mathbf{\}}`$ stands again for linear combination. In Eqs. (56)โ(59) we use the standard notation $`H_{s_l}^{(1,2)}(z)`$ for the Hankel functions, and $`I_{s_l}(z)`$ and $`K_{s_l}(z)`$ for the modified Bessel functions of the first and second kinds, respectively .
Let us first look at the weak-coupling regime. For bound states, the requirement that the solution be finite at infinity implies the rejection of the component $`I_{s_l}(\kappa r)`$ in Eq. (56), whereas the boundary condition at the origin, Eq. (A22), leads to the elimination of the component $`K_{s_l}(\kappa r)`$. In other words, there exist no bound states for a weak coupling. Physically, this state of affairs is reminiscent of the nonexistence of bound states for a repulsive potential; in fact, the combination of the potential itself and the centrifugal potential is effectively โrepulsiveโ when $`\lambda <\lambda _l^{()}`$. On the other hand, such a potential can produce nontrivial scattering, which, according to Eq. (57), is described by the solution
$`u_l(r)`$ $`=`$ $`\sqrt{r}\left[\stackrel{~}{A}_l^{(+)}H_{s_l}^{(1)}(kr)+\stackrel{~}{A}_l^{()}H_{s_l}^{(2)}(kr)\right]`$ (60)
$`\stackrel{(r\mathrm{})}{}`$ $`\sqrt{r}[\stackrel{~}{A}_l^{(+)}H_{l+\nu _0}^{(1)}\mathbf{(}\text{ }kr+(l+\nu _0s_l){\displaystyle \frac{\pi }{2}}\mathbf{)}\text{ }`$ (62)
$`+\stackrel{~}{A}_l^{()}H_{l+\nu _0}^{(2)}\mathbf{(}\text{ }kr+(l+\nu _0s_l){\displaystyle \frac{\pi }{2}}\mathbf{)}\text{ }],`$
where $`\stackrel{~}{A}_l^{(+)}=\stackrel{~}{A}_l^{()}`$ and the asymptotic behavior determines the energy-independent phase shifts
$$\delta _l^{(D_0)}=\left[\left(l+\nu _0\right)\sqrt{\left(l+\nu _0\right)^2\lambda }\right]\frac{\pi }{2}.$$
(63)
Equation (63) manifestly displays the scale invariance of the theory, generalizing the well-known three-dimensional results . Notice that the scattering matrix $`S_l^{(D_0)}=\mathrm{exp}\left[2i\delta _l^{(D_0)}\right]`$ has no poles, which is in agreement with the absence of bound states. Parenthetically, our conclusions relied on the choice of the boundary condition (A22) at the origin; the validity of this procedure for the inverse square potential will be proved in the next section.
Let us now consider the strong-coupling regime. For bound states, the requirement that the solution be finite at infinity implies again the rejection of the component $`I_{i\mathrm{\Theta }_l}(\kappa r)`$ in Eq. (58). However, the boundary condition at the origin, Eq. (A22), can neither eliminate the component $`K_{s_l}(\kappa r)`$ nor restrict the possible values of the energy. In effect, from Eq. (20) and elementary properties of the gamma function , the small-argument behavior of the modified Bessel function of the second kind and imaginary index is
$$K_{i\mathrm{\Theta }_l}(\kappa r)\stackrel{(r0)}{}\sqrt{\frac{\pi }{\mathrm{\Theta }_l\mathrm{sinh}\left(\pi \mathrm{\Theta }_l\right)}}\mathrm{sin}\left[\mathrm{\Theta }_l\mathrm{ln}\left(\frac{\kappa r}{2}\right)\delta _{\mathrm{\Theta }_l}\right]\left[1+O\left(r^2\right)\right],$$
(64)
where
$$\delta _{\mathrm{\Theta }_l}=\mathrm{}\left[\mathrm{ln}\mathrm{\Gamma }(1+i\mathrm{\Theta }_l)\right]$$
(65)
is the phase of $`\mathrm{\Gamma }(1+i\mathrm{\Theta }_l)`$. Then if Eq. (64) were a bound-state wave function, it would display an oscillatory behavior with an ever increasing frequency as $`r0`$. In particular, starting with any finite value of $`r`$, the function defined by Eq. (64) has an infinite number of zeros, at the points $`r_n=2\mathrm{exp}\left[\left(\delta _{\mathrm{\Theta }_l}n\pi \right)/\mathrm{\Theta }_l\right]/\kappa `$ (with $`n`$ integer), whence it represents a state lying above infinitely many bound states. In other words, the boundary condition has become ineffective as a screening tool, a situation that we may describe as a โlossโ of the boundary condition, and which permits the existence of a continuum of bound states extending from $`E=\mathrm{}`$ to $`E=0`$, in agreement with the conclusions of Ref. . In particular, the potential has no ground state, so that the Hamiltonian is no longer bounded from below and has lost its self-adjoint character . After due reflection, this situation makes sense: the potential has overcome the centrifugal barrier, a phenomenon that corresponds to the classical โfall of the particle to the centerโ . With regard to the scattering in this strong-coupling regime, due to the loss of the boundary condition, it is impossible to determine the phase of the wave function or relative values of the coefficients in Eq. (59); in other words, the scattering parameters are ill-defined.
In summary, for $`\lambda <\lambda _l^{()}`$, the inverse square potential is incapable of producing bound states but it scatters in a manifestly scale-invariant way; while for $`\lambda >\lambda _l^{()}`$, it destroys the discrete character of the bound-state spectrum, the uniqueness of the scattering solutions, and the self-adjoint nature of the Hamiltonian. Here we recognize some of the familiar features of unregularized transmuting potentials.
The failure of the inverse square potential to provide a discriminating boundary condition at the origin for the strong-coupling regime is now seen as the source of the singular behavior required by dimensional transmutation. In some sense, the key to our regularization procedure will be the restoration of a sensible boundary condition for $`r=0`$.
### B Loss of the Boundary Condition at the Origin for the Inverse Square Potential
Let us now reexamine the boundary condition at the origin, which seems to be the most important ingredient in the analysis of Subsection III A. For the inverse square potential, the standard argument leading to Eq. (A22) should be modified with the replacement (54), so that
$$R_l(r)=\frac{u_l(r)}{r^{(D_01)/2}}\stackrel{(r0)}{}\mathbf{\{}\text{ }r^{\alpha _{+,l}}\text{,}\text{ }r^{\alpha _{,l}}\mathbf{\}}\text{ },$$
(66)
where
$$\alpha _{\pm ,l}=\nu _0\pm s_l$$
(67)
are the exponents arising from the associated indicial equation, with $`s_l`$ given by Eq. (52). Obviously, Eq. (66) is the small-argument limit of Eqs. (56)-(59). However, the power-law functions (66) are no longer solutions of Laplaceโs equation, so that the rejection of the second solution, in the analysis leading to Eq. (A22), is not justified. In order to clarify this issue it proves convenient to consider a truncated potential
$$V_a(r)=\{\begin{array}{cc}\lambda /a^2\hfill & \text{for }r<a\hfill \\ \lambda /r^2\hfill & \text{for }ra\hfill \end{array},$$
(68)
which clearly satisfies $`V(r)=lim_{a0}V_a(r)`$. The truncated potential permits the use of regular boundary conditions for finite $`a`$, as a means of deducing the limiting behavior when $`a0`$. The solution in the presence of the truncated potential (68), for $`r`$ sufficiently small, is
$$R_l(r)\stackrel{(r0,a0)}{}\{\begin{array}{cc}C_l^{(+)}r^{\alpha _{+,l}}+C_l^{()}r^{\alpha _{,l}}\hfill & \mathrm{for}r>a\hfill \\ B_lr^l\hfill & \mathrm{for}r<a\hfill \end{array},$$
(69)
where the exponents $`\alpha _{\pm ,l}`$ are given by Eq. (67). Continuity of the logarithmic derivative for $`r=a`$, in the limit $`a0`$, provides the condition
$$\mathrm{\Omega }_l(a)=\frac{C_l^{(+)}}{C_l^{()}}=\frac{\sqrt{\lambda _l^{()}}+\sqrt{\lambda _l^{()}\lambda }}{\sqrt{\lambda _l^{()}}\sqrt{\lambda _l^{()}\lambda }}a^{2\sqrt{\lambda _l^{()}\lambda }}.$$
(70)
For the weak-coupling regime, $`\lambda <\lambda _l^{()}`$, the ratio (70) goes to infinity as $`a0`$, and the boundary condition becomes
$$R_l(r)\stackrel{(r0)}{}C_l^{(+)}r^{\alpha _{+,l}},$$
(71)
so that
$$u_l(r)\stackrel{(r0)}{}C_l^{(+)}\sqrt{r}r^{s_l}.$$
(72)
In other words, the choice is made in favor of the least divergent solution; then Eq. (72) implies that the boundary condition for the function $`u_l(r)`$ is (A22), just as for regular potentials. This confirms, a posteriori, that (A22) is a valid condition to apply in the weak-coupling case, as was assumed in the analysis of Subsection III A, which led to the proper selection of Bessel functions in Eqs. (56) and (57).
However, for the strong-coupling regime, $`\lambda >\lambda _l^{()}`$, the exponents $`\alpha _\pm `$ are complex conjugate values and both solutions have the same behavior. Then, no criterion can be given to discriminate the โgoodโ from the โbadโ solutions, so that the boundary condition at the origin is โlost.โ Correspondingly, the ratio of Eq. (70) has no definite limit for $`a0`$, displaying the same oscillatory behavior found in Subsection III A, i.e.,
$$\mathrm{\Omega }_l(a)\mathrm{exp}\left(2i\mathrm{\Theta }_l\mathrm{ln}a\right),$$
(73)
Then, as $`a0`$, the general solution reduces to the oscillatory form (64). Notice that Eq. (A22) is still satisfied, but it is not a boundary condition, as it has lost its discriminating power: both solutions are now allowed (cf. Eq. (66)).
In short, focusing on the behavior near the origin exclusively, we have clarified the meaning of the loss of boundary condition and rediscovered the basic features that had already been predicted in Subsection III A.
## IV INVERSE SQUARE POTENTIAL: RENORMALIZATION
In the first detailed treatment of the inverse square potential, Ref. , it was proposed that an additional orthogonality constraint be imposed on the eigenfunctions (58) in the strong-coupling regime. Effectively, this procedure restores the discrete nature of the spectrumโa direct application of Eq. (64) gives the energies $`E_{n^{}l}/E_{nl}=e^{2\pi (n^{}n)/\mathrm{\Theta }_l}`$. Nonetheless, the allowed bound-state levels still extend to $`\mathrm{}`$, i.e., the Hamiltonian remains unbounded from below. Subsequently, a number of different regularization techniques were introduced by properly adding regularization parameters of various kinds. It was soon realized that the strong-coupling Hamiltonian of the inverse square potential fails to be self-adjoint, despite it being a symmetric operator; as its deficiency indices are $`(1,1)`$, its solutions can be regularized with a single parameter in the form of self-adjoint extensions . Alternative attempts simply abandoned self-adjointness in favor of other requirements or interpretations; for example, in Ref. , the picture of the โfall of the particle to the centerโ is explicitly implemented by a non-Hermitian condition.
In our approach, we will follow the traditional path of enforcing the self-adjoint nature of the Hamiltonianโthis time using field-theoretic methods. Recently, a renormalized solution was presented along these lines , which replaced the orthogonality criterion of Ref. by a regular boundary condition at a cutoff point near the origin. However, this proposal was just limited to the one-dimensional case and only relied on real-space cutoff regularization. A second step along this path was the formulation of the $`D`$-dimensional generalization of the cutoff-regularization approach . In this section, we extend the results of Refs. using dimensional regularization and emphasize those features associated with dimensional transmutation.
### A Dimensional Regularization of the Inverse Square Potential
Dimensional regularization of the inverse square potential is done as follows. As discussed in Ref. , dimensional analysis implies that the $`D`$-dimensional regularization of the inverse square potential (Eqs. (1) and (3)) is of the homogeneous form
$$W^{(D)}(r)r^{(2ฯต)}.$$
(74)
We would like now to confirm this prediction and find the proportionality coefficient $`๐ฅ(ฯต)`$. First, we should find the momentum-space expression by means of a $`D_0`$-dimensional Fourier transform,
$$\stackrel{~}{W}(๐ค)=_{(D_0)}\left\{W(๐ซ)\right\}=d^{D_0}re^{i๐ค๐ซ}\frac{1}{r^2},$$
(75)
which we can be computed by means of Bochnerโs theorem , i.e.,
$$\stackrel{~}{W}(๐ค)=\frac{(2\pi )^{D_0/2}}{k^{D_0/21}}_0^{\mathrm{}}๐rr^{D_0/22}J_{D_0/21}(kr).$$
(76)
Applying the identity
$$_0^{\mathrm{}}x^\alpha J_\beta (az)๐z=2^\alpha a^{\alpha 1}\frac{\mathrm{\Gamma }(1/2+\beta /2+\alpha /2)}{\mathrm{\Gamma }(1/2+\beta /2\alpha /2)},$$
(77)
which is valid for $`\mathrm{Re}(\beta )1<\mathrm{Re}(\alpha )<1/2`$, the required transform is found to be
$$\stackrel{~}{W}(๐ค)=\frac{(4\pi )^{D_0/2}}{4k^{D_02}}\mathrm{\Gamma }(D_0/21).$$
(78)
Equation (78) seems to be restricted to $`2<\mathrm{Re}(D_0)<5`$; however, the right-hand side is a meromorphic function with poles at $`D_0=2,0,2,\mathrm{}`$ and is the desired analytic continuation of the integral. As we will see next, when the inverse Fourier transform is applied, the final analytically continued expression will have no such restriction. In effect, applying the dimensional-continuation prescription of Ref. to $`W^{(2)}(๐ซ)=1/r^2`$, we get
$`W^{(D)}(๐ซ_D)`$ $`=`$ $`{\displaystyle \frac{d^Dk_D}{(2\pi )^D}e^{i๐ค_D๐ซ_D}\left[\stackrel{~}{W}(๐ค_{๐_\mathrm{๐}})\right]_{๐ค_{D_0}๐ค_D}},`$ (79)
$`=`$ $`{\displaystyle \frac{(4\pi )^{D_0/2}}{4}}\mathrm{\Gamma }\left(D_0/21\right){\displaystyle \frac{d^Dk_D}{(2\pi )^D}e^{i๐ค_D๐ซ_D}\left[k_{D_0}^{D_0+2}\right]_{๐ค_{D_0}๐ค_D}}`$ (80)
$`=`$ $`{\displaystyle \frac{(2\pi )^{ฯต/2}}{r^{D/21}}}\mathrm{\hspace{0.17em}2}^{D_0/22}\mathrm{\Gamma }\left(D_0/21\right){\displaystyle _0^{\mathrm{}}}๐kk^{D/2D_0+2}J_{D/21}(kr),`$ (81)
which, by means of Eq. (77), amounts to
$$W^{(D_0ฯต)}(r)=\pi ^{ฯต/2}\mathrm{\Gamma }\left(1\frac{ฯต}{2}\right)r^{(2ฯต)}.$$
(82)
Even though the last integration is restricted to $`3D_0<\mathrm{Re}(D_0D)<2`$, analytic continuation of the right-hand of Eq. (82) allows us to extend its validity for arbitrary values of $`ฯต2,4,\mathrm{}`$ (notice that the previous restriction has been lifted); this is perfectly fine anyway, because all that is needed is the limit $`ฯต=0^+`$.
Then, for the dimensionally regularized problem, using the notation
$$๐ฅ(ฯต)=\pi ^{ฯต/2}\mathrm{\Gamma }\left(1\frac{ฯต}{2}\right),$$
(83)
we have the $`D`$-dimensional Schrรถdinger equation
$$\left[_D^2+E+\frac{\lambda \mu ^ฯต๐ฅ(ฯต)}{r^{2ฯต}}\right]\mathrm{\Psi }(๐ซ)=0,$$
(84)
which, for the state of angular momentum $`l`$, reads
$$\left\{\frac{d^2}{dr^2}+E+\frac{\lambda \mu ^ฯต๐ฅ(ฯต)}{r^{2ฯต}}\frac{\left[l+D(ฯต)/21\right]^21/4}{r^2}\right\}u_l(r)=0.$$
(85)
We are now ready to solve Eq. (85). As it stands, it does not look familiar when $`ฯต`$ is not an integer. However, it can be conveniently transformed into an asymptotically soluble problem by means of a duality transformation , whose properties we study in Appendix C. For the particular case at hand, applying the transformation (cf. (C22))
$$\{\begin{array}{c}\left|E\right|^{1/2}r=z^{2/ฯต}\hfill \\ |E|^{D/4}u(r)=w(z)z^{1/ฯต1/2}\hfill \end{array},$$
(86)
the dual of Eq. (85) becomes Eq. (C26), which we write in the condensed form
$$\left\{\frac{d^2}{dz^2}+\stackrel{~}{\eta }\stackrel{~}{๐ฑ}_ฯต(z)\frac{[p_l(ฯต)]^21/4}{z^2}\right\}w_{l,ฯต}(z)=0,$$
(87)
with a dual dimensionless energy
$$\stackrel{~}{\eta }=\frac{4\lambda \mu ^ฯต๐ฅ(ฯต)}{ฯต^2}|E|^{ฯต/2},$$
(88)
a dual potential energy term,
$$\stackrel{~}{๐ฑ}_ฯต(z)=\sigma \frac{4}{ฯต^2}z^{4/ฯต2},$$
(89)
and a dual angular momentum variable
$$p_l(ฯต)=\frac{2}{ฯต}\left[l+\frac{D(ฯต)}{2}1\right].$$
(90)
In our subsequent analysis, we will often rewrite Eq. (90) in the form
$$p=p_l(ฯต)=\frac{2}{ฯต}\left(\lambda _l^{()}\right)^{1/2}1=\frac{2}{ฯต}\left(\lambda _l^{()}\right)^{1/2}\left[1\frac{ฯต}{2}\left(\lambda _l^{()}\right)^{1/2}\right],$$
(91)
in terms of the critical coupling (53). Equation (91) provides an alternative expansion parameter $`p`$, such that, as $`ฯต0`$, $`p\mathrm{}`$. It should be noticed in passing that Eq. (88) already provides the functional form (4) required for any possible eigenvalue equation of the inverse square potential, with
$$\mathrm{\Xi }(ฯต)=\frac{4๐ฅ(ฯต)}{ฯต^2}\stackrel{~}{\eta }^1.$$
(92)
As usual, as the problem is now regularized, Eq. (87) should be solved with the regular boundary condition at the origin, Eq. (A22), so that
$$w_{l,ฯต}(0)=0.$$
(93)
Even though the solutions to Eq. (87) cannot be expressed in terms of any standard special functions for arbitrary $`ฯต>0`$, it is still possible to write their asymptotic form with respect to $`ฯต0`$, in terms of Bessel functions, as we show below. The key to this asymptotic analysis lies in the singular limiting form of the transformed potential (89), namely,
$$\underset{ฯต0}{lim}\stackrel{~}{๐ฑ}_ฯต(z)=\{\begin{array}{cc}0\hfill & \mathrm{for}z<1\hfill \\ \sigma \mathrm{}\hfill & \mathrm{for}z>1\hfill \end{array};$$
(94)
in particular, for $`E<0`$ (i.e., $`\sigma =1`$), $`\stackrel{~}{๐ฑ}_ฯต(z)`$ behaves as an infinite hyperspherical potential well. Because of the singular nature of the limit (94), an asymptotically exact hierarchy emerges, whereby Eq. (87) is split into two distinct differential equations: the first one, without the term $`\stackrel{~}{๐ฑ}_ฯต(z)`$, for the interior region $`z<z_21`$,
$$\left\{\frac{d^2}{dz^2}+\stackrel{~}{\eta }\frac{[p_l(ฯต)]^21/4}{z^2}\right\}w_{l,ฯต}^{(<)}(z)=0;$$
(95)
and the second one, without the terms proportional to $`1/z^2`$, for the exterior region $`z>z_21`$,
$$\left[\frac{d^2}{dz^2}\stackrel{~}{๐ฑ}_ฯต(z)\right]w_{l,ฯต}^{(>)}(z)=0.$$
(96)
Equation (95) is a Bessel differential equation of order $`p=p_l(ฯต)`$, whose regular solutions that satisfy the boundary condition (93) are of the form
$$w_{l,ฯต}^{(<)}(z)=B_l\sqrt{z}J_p(\stackrel{~}{\eta }^{1/2}z).$$
(97)
On the other hand, Eq. (96) can be taken to a standard Bessel form by another duality transformation; it is easy to verify that
$$w_{l,ฯต}^{(>)}(z)=\sqrt{z}๐_{ฯต/4}(\sqrt{\sigma }z^{2/ฯต}),$$
(98)
where $`๐_{ฯต/4}`$ stands for a linear combination of ordinary Bessel functions of order $`ฯต/4`$; in fact, more explicitly,
$$w_{l,ฯต}^{(>)}(z)=\{\begin{array}{cc}A_l\sqrt{z}K_{ฯต/4}(z^{2/ฯต})\hfill & \mathrm{for}\sigma =1\hfill \\ \sqrt{z}\left[\stackrel{~}{A}_l^{(+)}H_{ฯต/4}^{(1)}(z^{2/ฯต})+\stackrel{~}{A}_l^{()}H_{ฯต/4}^{(2)}(z^{2/ฯต})\right]\hfill & \mathrm{for}\sigma =1\hfill \end{array},$$
(99)
where the first line includes a modified Bessel function of the second kind, whereas the second line includes a linear combination of Hankel functions. Their physical meaning becomes transparent when transformed back from the dual to the original space, in which
$$u_l(r)\stackrel{(r\mathrm{},ฯต0)}{}\{\begin{array}{cc}A_l\sqrt{r}K_0(\kappa r)\hfill & \mathrm{for}E=\kappa ^2<0\hfill \\ \sqrt{r}\left[\stackrel{~}{A}_l^{(+)}H_0^{(1)}(kr)+\stackrel{~}{A}_l^{()}H_0^{(2)}(kr)\right]\hfill & \mathrm{for}E=k^2>0\hfill \end{array}.$$
(100)
Equation (100) represents the asymptotic behavior of the wave function, when $`r`$ is large, for the bound-state ($`E<0`$) and scattering ($`E>0`$) sectors of the inverse square potential. Of course, the asymptotic behavior of Eq. (100) justifies a posteriori the choice of Bessel functions in Eq. (99).
The boundary point $`z_2`$ separating the two regimes described by Eqs. (95) and (96) is infinitesimally close to $`z=1`$, but further precision is required. In effect, the critical separation takes place at the point $`z_2`$ where
$$\left|\stackrel{~}{\eta }\frac{[p_l(ฯต)]^21/4}{z_2^2}\right|=\left|\stackrel{~}{๐ฑ}_ฯต(z_2)\right|,$$
(101)
because, away from it, the term $`\stackrel{~}{๐ฑ}_ฯต(z_2)`$ will abruptly become negligible for $`z<z_2`$ and will overwhelmingly dominate over the other terms for $`z>z_2`$. From Eqs. (88)โ(91), it follows that the condition (101) amounts to
$$p^2\left(\frac{\lambda }{\lambda _l^{()}}1\right)\left[1+O\left(\frac{1}{p}\right)\right]=\frac{p^2}{\lambda _l^{()}}z_2^{4/ฯต}\left[1+O\left(\frac{1}{p}\right)\right],$$
(102)
so that
$$z_2=\mathrm{\Theta }_l^{ฯต/2},$$
(103)
where $`\mathrm{\Theta }_l`$ is given by Eq. (55). With the value of $`z_2`$ so determined, the solutions to Eqs. (95) and (96) should be matched at $`z=z_2`$ through the logarithmic derivative
$$(\mathrm{\Theta }_l)=z\frac{d\mathrm{ln}\left(w/\sqrt{z}\right)}{dz}|_{z=z_2}.$$
(104)
For the interior solution (97), the logarithmic derivative takes the form
$$^{(<)}(\mathrm{\Theta }_l)=y\frac{d\mathrm{ln}J_p(y)}{dy}|_{y=\stackrel{~}{\eta }^{1/2}z_2},$$
(105)
whereas, for the exterior solution (97), it becomes
$$^{(>)}(\mathrm{\Theta }_l)=\frac{2}{ฯต}\mathrm{\Theta }_l\frac{d\mathrm{ln}๐_{ฯต/4}(\mathrm{\Theta }_l)}{d\mathrm{\Theta }_l}.$$
(106)
Having developed the general framework for the analysis of the inverse square potential in terms of its dual problem, we now turn to the distinct details of the bound-state and scattering sectors.
### B Bound-State Sector for the Inverse Square Potential
The bound states of the inverse square potential (if any) should be characterized by the energy condition $`E<0`$, which essentially converts the dual potential into an infinite hyperspherical well in the limit $`ฯต0^+`$, as displayed by Eq. (94). This suggests that, as a first approximation, the boundary condition at $`z_2`$ may be replaced by
$$w_{l,ฯต}^{(<)}(1)0,$$
(107)
whence Eq. (97) immediately gives the spectrum through the condition
$$J_p(\stackrel{~}{\eta }^{1/2})0,$$
(108)
from which the corresponding eigenvalues are
$$\stackrel{~}{\eta }_{n_r}\left(j_{p,n_r}\right)^2,$$
(109)
where $`j_{p,n_r}`$ is the $`n_r`$th zero of the Bessel function of order $`p`$. Therefore, from Eq. (88), the regularized eigenvalue condition becomes
$$\lambda \left[\frac{4๐ฅ(ฯต)(j_{p,n_r})^2}{ฯต^2}\right]\left(\frac{|E|}{\mu ^2}\right)^{ฯต/2}1,$$
(110)
where $`n_r`$ is now confirmed as the radial quantum numberโthe ordinal number for the stationary radial wave functions. Equation (110) is of the form of the master eigenvalue equation (4), with an energy generating function
$$\mathrm{\Xi }_{n_rl}(ฯต)\frac{4๐ฅ(ฯต)}{ฯต^2}\left(j_{p,n_r}\right)^2.$$
(111)
Even though the remarks of the previous paragraph are essentially correct, for a full comparison with the scattering sector of the theory, it is necessary to evaluate the logarithmic derivatives (105) and (106), which will give extra terms in the energy expressions. Therefore, we need to first analyze the asymptotic behavior of the interior solution (97) with respect to $`p\mathrm{}`$, and then find the proper values of the logarithmic derivatives.
The limit $`ฯต0`$ of Eq. (97) relies on well-known properties of the Bessel functions of large order . We will start with Debyeโs asymptotic expansion,
$$J_p(p\mathrm{sec}\beta )\stackrel{(p\mathrm{})}{}\sqrt{\frac{2}{\pi p\mathrm{tan}\beta }}\left\{\mathrm{cos}\left[p\mathrm{tan}\beta p\beta \frac{\pi }{4}\right]+O(p^1)\right\},$$
(112)
where the argument will become
$$y=p\mathrm{sec}\beta =\stackrel{~}{\eta }^{1/2}z_2.$$
(113)
For our analysis, a few relations between these variables are in order; in particular,
$$y=p\sqrt{\frac{\lambda }{\lambda _l^{()}}}\left[1+O\left(ฯต\right)\right]=p\sqrt{1+\frac{\mathrm{\Theta }_l^2}{\lambda _l^{()}}}\left[1+O\left(ฯต\right)\right],$$
(114)
and
$$\mathrm{tan}\beta =\sqrt{\left(y/p\right)^21}=\frac{\mathrm{\Theta }_l}{\left(\lambda _l^{()}\right)^{1/2}}.$$
(115)
Then the logarithmic derivative (105) becomes
$$^{(<)}(\mathrm{\Theta }_l)=p\frac{\mathrm{\Theta }_l}{\left(\lambda _l^{()}\right)^{1/2}}\mathrm{tan}\left\{p\left[\frac{\mathrm{\Theta }_l}{\left(\lambda _l^{()}\right)^{1/2}}\mathrm{tan}^1\left(\frac{\mathrm{\Theta }_l}{\left(\lambda _l^{()}\right)^{1/2}}\right)\right]\frac{\pi }{4}\right\}+O(1).$$
(116)
Equation (116) displays an anomalous behavior as $`p\mathrm{}`$ in that the tangent function will oscillate wildly from $`\mathrm{}`$ to $`\mathrm{}`$, thus rendering the regularized problem ill-defined. The cure for this behavior is afforded by the renormalization of the coupling constant $`\lambda =\lambda (ฯต)`$, which implies a corresponding renormalization of $`\mathrm{\Theta }_l=\mathrm{\Theta }_l(ฯต)`$ (from Eq. (55)), in such a way that $`\mathrm{\Theta }_l0`$ as $`ฯต0`$. Then this amounts to the realization of the limiting critical coupling $`\lambda \lambda _l^{()}+0^+`$. A more careful analysis of this limiting behavior is afforded by
$$=\mathrm{tan}\beta \beta =\frac{\beta ^3}{3}+O(\beta ^5)=\frac{\mathrm{\Theta }_l^3}{3\left(\lambda _l^{()}\right)^{3/2}}+O(\mathrm{\Theta }_l^5),$$
(117)
so that the finiteness of the argument of the tangent implies that $`p\mathrm{\Theta }_l^3`$ should be finite; then
$$\mathrm{\Theta }_l=O\left(p^{1/3}\right)$$
(118)
and
$$y=p+p^{1/3}x,$$
(119)
where $`x`$ is a finite variable.
On the other hand, using the small-argument behavior of the modified Bessel function of the second kind , Eq. (20), it follows that
$$K_{ฯต/4}(z^{2/ฯต})\stackrel{(z0)}{}\left[\frac{2}{ฯต}\left(\frac{1}{\sqrt{z}}\sqrt{z}\right)\frac{1}{2}\left(\gamma \mathrm{ln}2\right)\left(\frac{1}{\sqrt{z}}+\sqrt{z}\right)\right]\left[1+O(ฯต^2,z^{4/ฯต})\right],$$
(120)
so that, from Eqs. (103), (106), and (118), the logarithmic derivative becomes
$$^{(>)}(\mathrm{\Theta }_l)=\frac{2}{ฯต}\left(\mathrm{ln}\mathrm{\Theta }_lc\right)^1\left[1+O(ฯต^{2/3})\right],$$
(121)
where
$$c=\mathrm{ln}2\gamma .$$
(122)
Equation (121) implies that, as $`ฯต0`$,
$$\frac{w^{(>)}(z_2)}{\sqrt{z_2}}\stackrel{(ฯต0)}{}\frac{ฯต}{2}\left(\mathrm{ln}\mathrm{\Theta }_lc\right)\frac{d}{dz}\left[\frac{w^{(>)}(z)}{\sqrt{z}}\right]|_{z=z_2},$$
(123)
so that continuity of the wave function implies that Eq. (107) is indeed correct to zeroth order.
The interior logarithmic derivative can now be evaluated from the asymptotic expansion
$$J_p\left(p+p^{1/3}x\right)\stackrel{(p\mathrm{})}{}\frac{2^{1/3}}{p^{1/3}}Ai\left(2^{1/3}x\right)+O\left(p^1\right),$$
(124)
where $`Ai`$ is the Airy function of the first kind . The zeroth order approximation to the logarithmic derivative amounts to $`w(1)=0`$ or $`^{(<)}=\mathrm{}`$, so that the value of $`x`$, to this order, is provided by the zeros of the Airy function,
$$Ai\left(2^{1/3}x_{n_r}^{(0)}\right)=0.$$
(125)
More precisely, if $`a_{n_r}`$ is the $`n_r`$th negative zero of $`Ai`$, then
$$C_{n_r}=x_{n_r}^{(0)}=2^{1/3}a_{n_r},$$
(126)
where we have identified the leading contribution in the asymptotic formula for the zeros of the Bessel function of large order ,
$$j_{p,n_r}=p+C_{n_r}p^{1/3}+O\left(p^{2/3}\right);$$
(127)
for example, $`C_1=1.8558`$, $`C_2=3.2447`$, etc. It should be pointed out that, from the asymptotic form of the Airy functions, which amounts to a WKB integration of Eq. (95), an approximateโbut extremely accurateโformula can be derived for these coefficients, namely ,
$$C_{n_r}\frac{\left(3\pi \right)^{2/3}}{2}\left(n_r\frac{1}{4}\right)^{2/3}.$$
(128)
Then, for the next order,
$$x_{n_r}=C_{n_r}+\delta _{n_r},$$
(129)
with $`\delta _{n_r}C_{n_r}`$, one can expand the Airy function in a Taylor series in the neighborhood of $`x_{n_r}^{(0)}`$, so that Eq. (105) becomes
$`^{(<)}(\mathrm{\Theta }_l)`$ $`=`$ $`2^{1/3}p^{2/3}{\displaystyle \frac{Ai^{}\left(2^{1/3}x\right)}{Ai\left(2^{1/3}x\right)}}\left[1+O\left(p^{2/3}\right)\right]`$ (130)
$`=`$ $`{\displaystyle \frac{p^{2/3}}{\delta _{n_r}}}\left[1+O(\delta _{n_r},p^{2/3})\right].`$ (131)
The equality of the logarithmic derivatives, Eqs. (121) and (131), gives
$$\delta _{n_r}=p^{1/3}\left(\lambda _l^{()}\right)^{1/2}\left[c+\mathrm{ln}\mathrm{\Theta }_l\right]\left[1+O\left(ฯต^{1/3}\right)\right],$$
(132)
whence
$$y_{n_r}=p\left\{1+C_{n_r}p^{2/3}+\left(\lambda _l^{()}\right)^{1/2}\left[c+\mathrm{ln}\mathrm{\Theta }_l\right]p^1+O\left(p^{4/3}\right)\right\}.$$
(133)
Therefore, from Eqs. (103) and (113), the correct replacement for Eq. (109) is
$$\stackrel{~}{\eta }_{n_r}=\left(\frac{y}{z_2}\right)^2=p^2\left\{1+2C_{n_r}p^{2/3}2\left(\lambda _l^{()}\right)^{1/2}cp^1+O\left(p^{4/3}\right)\right\}.$$
(134)
Finally, from Eq. (83), the expansion of $`๐ฅ(ฯต)`$ about $`ฯต=0^+`$ is
$$๐ฅ(ฯต)=1+\frac{ฯต}{2}\left(\gamma +\mathrm{ln}\pi \right)+O(ฯต^2),$$
(135)
which combined with Eqs. (91) and (134), gives the asymptotically exact expression for the eigenvalue function of Eqs. (4) and (92), namely,
$`\mathrm{\Xi }_{n_rl}(ฯต)`$ $`=`$ $`{\displaystyle \frac{1}{\lambda _l^{()}}}\left\{12p^{2/3}C_{n_r}+p^1\left(\lambda _l^{()}\right)^{1/2}\left[\mathrm{ln}4\pi \gamma +2\left(\lambda _l^{()}\right)^{1/2}\right]+O\left(p^{4/3}\right)\right\}`$ (136)
$`=`$ $`{\displaystyle \frac{1}{\lambda _l^{()}}}\left\{12^{1/3}C_{n_r}(\lambda _l^{()})^{1/3}ฯต^{2/3}+{\displaystyle \frac{ฯต}{2}}\left[\mathrm{ln}4\pi \gamma +2(\lambda _l^{()})^{1/2}\right]+O\left(ฯต^{4/3}\right)\right\},`$ (137)
which is of the form
$$\mathrm{\Xi }_n(ฯต)=\left[L_n(ฯต)\right]^1\left[1+\frac{ฯต}{2}๐ข_n(ฯต)\right],$$
(138)
with a constant critical coupling function
$$L_{n_rl}=\lambda _l^{()}=\left(l+\frac{D_0}{2}1\right)^2$$
(139)
(independent of $`n_r`$), and with
$$๐ข_{n_rl}(ฯต)=2^{4/3}C_{n_r}(\lambda _l^{()})^{1/3}ฯต^{1/3}+\left[\mathrm{ln}4\pi \gamma +2(\lambda _l^{()})^{1/2}\right]+O(ฯต^{1/3}).$$
(140)
From Eq. (140), we see that
$$๐ข_{_{(\mathrm{gs})}}(ฯต)=2^{4/3}C_1(\lambda _{_{(\mathrm{gs})}}^{()})^{1/3}ฯต^{1/3}+\left[\mathrm{ln}4\pi \gamma +2(\lambda _{_{(\mathrm{gs})}}^{()})^{1/2}\right]+O(ฯต^{1/3}),$$
(141)
where
$$\lambda _{_{(\mathrm{gs})}}^{()}=\lambda _0^{()}=\nu _0^2=\left(D_0/21\right)^2$$
(142)
is the โprincipal coupling constant,โ i.e., the critical coupling for the ground state ($`l=0`$).
Equations (4) and (138)โ(140) provide the regularized energies
$`|E_{n_rl}|`$ $`=`$ $`\mu ^2\left({\displaystyle \frac{\lambda }{\lambda _l^{()}}}\right)^{2/ฯต}\mathrm{exp}\left[๐ข_{n_rl}(ฯต)\right]`$ (143)
$`=`$ $`\mu ^2\left({\displaystyle \frac{\lambda }{\lambda _l^{()}}}\right)^{2/ฯต}\mathrm{exp}\left\{2^{4/3}C_{n_r}\left(\lambda _l^{()}\right)^{1/3}ฯต^{1/3}+\left[\mathrm{ln}4\pi \gamma +2\left(\lambda _l^{()}\right)^{1/2}\right]\right\}.`$ (144)
Ground-state renormalization can be implemented from Eq. (144), by means of the regularized coupling
$$\lambda (ฯต)=\left[\mathrm{\Xi }_{_{(\mathrm{gs})}}(ฯต)\right]^1\left[1+\frac{ฯต}{2}g^{(0)}+o(ฯต)\right],$$
(145)
which now becomes
$$\lambda (ฯต)=\lambda _{_{(\mathrm{gs})}}^{()}\left\{1+2^{1/3}\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/3}ฯต^{2/3}C_1+\frac{ฯต}{2}\left[g^{(0)}\left(\mathrm{ln}4\pi \gamma \right)2(\lambda _{_{(\mathrm{gs})}}^{()})^{1/2}\right]\right\}+o(ฯต),$$
(146)
with an arbitrary finite part $`g^{(0)}`$, so that, for the ground state,
$$E_{_{(\mathrm{gs})}}=\mu ^2\mathrm{exp}\left[g^{(0)}\right]\mu ^2$$
(147)
(cf. Eq. (29)), while for any regularized bound state
$`|E_{n_rl}|=\mu ^2\left({\displaystyle \frac{\lambda _{_{(\mathrm{gs})}}}{\lambda _l^{()}}}\right)^{2/ฯต}\mathrm{exp}\{2^{4/3}[C_{n_r}\left(\lambda _l^{()}\right)^{1/3}`$ $``$ $`C_1\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/3}]ฯต^{1/3}`$ (148)
$`+`$ $`g^{(0)}+2[\left(\lambda _l^{()}\right)^{1/2}\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/2}]\}.`$ (149)
Equation (146) explicitly shows that
$$\lambda (ฯต)\stackrel{(ฯต0)}{}\lambda _{_{(\mathrm{gs})}}^{()}+0^+;$$
(150)
in other words, when the system is renormalized, the coupling becomes critically strong with respect to the ground state (which, as we will see below, has $`l=0`$).
Equations (137)โ(150) are easily interpreted. As expected, the critical coupling defined by Eq. (53) from the unregularized theory becomes the critical coupling for the regularized theory, Eq. (139), with respect to dimensional transmutation. We already know, from the criteria of the general theory of dimensional transmutation , that Eq. (137) implies the existence of a ground state alone. However, we will now illustrate the general arguments for this particular problem.
The value of the critical coupling depends on the angular momentum quantum number $`l`$ but is independent of the radial quantum number $`n_r`$. This, combined with the form of $`๐ข_{n_rl}(ฯต)`$ in Eq. (140), imposes very stringent conditions on the existence of bound states, as we shall see next.
The dependence of $`\lambda _l^{()}`$ with respect to $`l`$ requires that only $`l=0`$ states (if any) be allowed, as implied by the following argument. First, only a finite number of states can exist with different angular momentum numbers $`l`$. In effect, let us assume the existence of a given state with angular momentum $`l_0`$; as the coupling is critically strong, then $`\lambda =\left(l_0+\nu _0\right)^2`$, so that the potential is โweakโ and has no bound states for all $`l>l_0`$โfor weak coupling, the unregularized theory suffices. In other words, the only allowed states are those with $`0ll_0`$. Let us now see that $`ล_00`$ would lead to a contradiction; in effect, if $`ล_0>0`$, then a state with $`l<l_0`$ would potentially exist as it would correspond to the strong regime; however, as $`ll_0`$, the coupling would not be critical with respect to $`l`$, rendering the theory ill-defined. In other words, if the ground state exists, it should have $`l=0`$ (as expected) and all states with $`l>0`$ are forbidden.
Therefore, the ground state is characterized by the quantum numbers
$$(\mathrm{gs})\left(n_r=1,l=0\right),$$
(151)
and only hypothetical states with $`l=0`$ survive the renormalization process as bound states.
The next question is whether states with $`l=0`$ but $`n_r0`$ can survive renormalization. Of course, the ratio of Eqs. (147) and (149) provides the answer: starting with the ground-state energy, any other such state would have an energy overwhelmingly suppressed by an exponential factor,
$$\left|\frac{E_{n_r0}}{E_{_{(\mathrm{gs})}}}\right|=\mathrm{exp}\left[2^{4/3}\left(C_{n_r}C_1\right)\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/3}ฯต^{1/3}\right]\stackrel{(ฯต0,n_r>1)}{}0.$$
(152)
Furthermore, for these states, the wave function (100) has an ill-defined limit, so that they cease to exist when $`ฯต0`$. In short, the singular nature of the potential is responsible for the destruction of all candidates for a renormalized bound state, except for the limit of the ground state of the regularized theoryโwhich, upon renormalization, becomes the ground state of the system and acquires the finite energy value (147).
Finally, the ground-state wave function can also be explicitly derived. From Eqs. (100), (147), and (A15), it follows that the ground state wave function is given by
$$\mathrm{\Psi }_{_{(\mathrm{gs})}}(๐ซ)=\sqrt{\mathrm{\Gamma }(\nu +1)\left(\frac{\mu ^2}{\pi }\right)^{\nu +1}}\frac{K_0(\mu r)}{\left(\mu r\right)^\nu },$$
(153)
which is similar to that of the two-dimensional delta-function potential, with which it coincides when $`D_0=2`$.
### C Scattering Sector for the Inverse Square Potential
For the scattering sector of the theory we will match the interior and exterior solutions by means of the parameter
$$\omega (\mathrm{\Theta }_l)=\frac{ฯต}{2}(\mathrm{\Theta }_l)=\mathrm{\Theta }_l\frac{d\mathrm{ln}๐_{ฯต/4}(\mathrm{\Theta }_l)}{d\mathrm{\Theta }_l}.$$
(154)
In particular, the exterior scattering problem is described by the second line in Eqs. (99) and (100), so that Eq. (106) can now be rewritten as
$$\omega ^{(>)}(\mathrm{\Theta }_l)=\mathrm{\Theta }_l\frac{d\mathrm{ln}๐_{ฯต/4}(\mathrm{\Theta }_l)}{d\mathrm{\Theta }_l},$$
(155)
which is explicitly given by
$$\omega ^{(>)}(\mathrm{\Theta }_l)=\frac{\stackrel{~}{A}_l^{(+)}\mathrm{\Theta }_lH_{ฯต/4}^{(1)}(\mathrm{\Theta }_l)+\stackrel{~}{A}_l^{()}\mathrm{\Theta }_lH_{ฯต/4}^{(2)}(\mathrm{\Theta }_l)}{\stackrel{~}{A}_l^{(+)}H_{ฯต/4}^{(1)}(\mathrm{\Theta }_l)+\stackrel{~}{A}_l^{()}H_{ฯต/4}^{(2)}(\mathrm{\Theta }_l)}.$$
(156)
Equation (156) can be evaluated from the small-argument behavior of the Hankel functions ,
$`H_p^{(1,2)}(z)`$ $`\stackrel{(z0)}{}`$ $`\pm {\displaystyle \frac{e^{ip\pi /2}}{\pi i}}\left[e^{ip\pi /2}\mathrm{\Gamma }(p)\left({\displaystyle \frac{z}{2}}\right)^p+e^{\pm ip\pi /2}\mathrm{\Gamma }(p)\left({\displaystyle \frac{z}{2}}\right)^p\right]\left[1+O(z^2)\right],`$ (157)
so that
$$\omega _l=\omega ^{(>)}(\mathrm{\Theta }_l)=\frac{2i}{\pi }\frac{\stackrel{~}{A}_l^{(+)}\stackrel{~}{A}_l^{()}}{(1+iQ_l)\stackrel{~}{A}_l^{(+)}+(1iQ_l)\stackrel{~}{A}_l^{()}},$$
(158)
where
$$Q_l=\frac{2}{\pi }\left(c+\mathrm{ln}\mathrm{\Theta }_l\right),$$
(159)
with $`c`$ defined in Eq. (122). Then the scattering matrix can be derived from the asymptotic expansion of the Hankel functions as $`r\mathrm{}`$, which according to Eq. (100) is of the form
$$u(r)\stackrel{(r\mathrm{},ฯต0)}{}\sqrt{r}\left[A_l^{(+)}H_{l+\nu _0}^{(1)}\mathbf{(}\text{ }kr+\left(l+\nu _0\right)\frac{\pi }{2}\mathbf{)}\text{ }+A_l^{()}H_{l+\nu _0}^{(2)}(kr)\mathbf{(}\text{ }kr+\left(l+\nu _0\right)\frac{\pi }{2}\mathbf{)}\text{ }\right],$$
(160)
where the coefficients for the $`s`$ wave ($`l=0`$) are related via
$$A_0^{(\pm )}=\stackrel{~}{A}_0^{(\pm )}e^{\pm i(D_02)\pi /4};$$
(161)
then the $`s`$-wave scattering matrix elementsโfrom Eqs. (158), (161), and (B13)โare given by
$$S_0^{(D_0)}(k)=e^{i\nu _0\pi }\stackrel{~}{S}_0^{(D_0)}(k),$$
(162)
with $`\stackrel{~}{S}_0^{(D_0)}(k)=\stackrel{~}{A}_0^{(+)}/\stackrel{~}{A}_0^{()}`$ of the form
$$\stackrel{~}{S}_0^{(D_0)}(k)=\frac{1+i\left(2\omega _{_{(\mathrm{gs})}}^1/\pi Q_{_{(\mathrm{gs})}}\right)}{1i\left(2\omega _{_{(\mathrm{gs})}}^1/\pi Q_{_{(\mathrm{gs})}}\right)},$$
(163)
and where $`\omega _{_{(\mathrm{gs})}}=\omega _0`$ and $`Q_{_{(\mathrm{gs})}}=Q_0`$.
In order to obtain an explicit expression for the scattering matrix in terms of the energy, we need to evaluate the coefficient $`\omega _{_{(\mathrm{gs})}}`$ in Eq. (163). This program can be implemented through the expansion of the dual energy $`\stackrel{~}{\eta }`$, defined in Eq. (88), which should now be derived again for the scattering sector of the theory. With that purpose in mind, making use of Eqs. (83), (91), and
$$\left|\frac{E}{\mu ^2}\right|^{ฯต/2}=1\frac{ฯต}{2}\mathrm{ln}\left|\frac{E}{\mu ^2}\right|+O\left(ฯต^2\right),$$
(164)
the required expression becomes (for $`l=0`$)
$$\stackrel{~}{\eta }=p^2\frac{\lambda }{\lambda _{_{(\mathrm{gs})}}^{()}}\left\{1+\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/2}\left[\mathrm{ln}\pi +\gamma +2\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/2}\mathrm{ln}\left|\frac{E}{\mu ^2}\right|\right]p^1+O\left(p^2\right)\right\}.$$
(165)
In addition, we have already renormalized the bound-state sector of the theory, with the result that the coupling constant should behave as dictated by Eq. (146), whence
$$\stackrel{~}{\eta }^{1/2}=p\left\{1+C_1p^{2/3}+\frac{ฯต}{4}\left[2c+g^{(0)}\mathrm{ln}\left|\frac{E}{\mu ^2}\right|\right]+o(ฯต)\right\},$$
(166)
from which it follows that the variable $`y=\stackrel{~}{\eta }^{1/2}z_2`$, defined by Eq. (113), is of the form (119), with
$$x=C_1+\left[\left(c+\mathrm{ln}\mathrm{\Theta }_{_{(\mathrm{gs})}}\right)\frac{1}{2}\mathrm{ln}\left|\frac{E}{E_{_{(\mathrm{gs})}}}\right|\right]\left(\lambda _{_{(\mathrm{gs})}}^{()}\right)^{1/2}p^{1/3},$$
(167)
and $`c`$ defined in Eq. (122). Then the logarithmic derivative (105) can be derived for the scattering sector using an argument similar to the one employed in Eq. (131), i.e., with $`\delta =xC_1`$,
$`^{(<)}(\mathrm{\Theta }_l)`$ $`=`$ $`2^{1/3}p^{2/3}{\displaystyle \frac{Ai^{}\left(2^{1/3}x\right)}{Ai\left(2^{1/3}x\right)}}\left[1+O\left(p^{2/3}\right)\right]`$ (168)
$`=`$ $`{\displaystyle \frac{p^{2/3}}{xC_1}}\left[1+O\left(p^{2/3}\right)\right].`$ (169)
As a consequence, we conclude that the parameter defined in Eq. (154) becomes
$$\omega _{_{(\mathrm{gs})}}=\omega ^{(<)}(\mathrm{\Theta }_{_{(\mathrm{gs})}})=\left[\left(c+\mathrm{ln}\mathrm{\Theta }_{_{(\mathrm{gs})}}\right)\frac{1}{2}\mathrm{ln}\left|\frac{E}{E_{_{(\mathrm{gs})}}}\right|\right]^1.$$
(170)
Finally, from Eqs. (159), (162), (163), and (170), the S-matrix reads
$$S_0^{(D_0)}(k)=e^{i\nu _0\pi }\frac{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)+i\pi }{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)i\pi },$$
(171)
while the phase shifts are implicitly given by the expression
$$\mathrm{tan}\mathbf{(}\text{ }\delta _0^{(D_0)}(k)\frac{\pi \nu _0}{2}\mathbf{)}\text{ }=\frac{\pi }{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)},$$
(172)
and the partial scattering amplitude reads
$$a_0^{(D_0)}(k)=\frac{1}{2ik^{(D_01)/2}}\frac{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)\left(e^{i\nu _0\pi }1\right)+i\pi \left(e^{i\nu _0\pi }+1\right)}{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)i\pi }.$$
(173)
Equations (171), (172), and (173) are remarkably similar to Eqs. (43), (44), and (46) for the two-dimensional delta function potential. Table I summarizes the main results of the scattering for an arbitrary number of dimensions $`D_0`$.
###### TABLE I
. Phase Shift $`\delta _0^{(D_0)}(k)`$, Scattering Matrix $`S_0^{(D_0)}(k)`$, and Partial Scattering Amplitude $`a_0^{(D_0)}(k)`$ for the Inverse Square Potential, as a Function of the Geometric Dimension $`D_0`$ of Position Space.
| $`D_0`$ | $`\mathrm{tan}\delta _0^{(D_0)}(k)`$ | $`S_0^{(D_0)}(k)`$ | $`a_0^{(D_0)}(k)k^{(D_01)/2}`$ |
| --- | --- | --- | --- |
| $`1(mod4)`$ | $`{\displaystyle \frac{\pi L}{\pi +L}}`$ | $`{\displaystyle \frac{\left(1+i\right)\pi +\left(1i\right)L}{\left(1i\right)\pi +\left(1+i\right)L}}`$ | $`{\displaystyle \frac{\pi L}{\left(1i\right)\pi +\left(1+i\right)L}}`$ |
| $`2(mod4)`$ | $`{\displaystyle \frac{\pi }{L}}`$ | $`{\displaystyle \frac{L+i\pi }{Li\pi }}`$ | $`{\displaystyle \frac{\pi }{Li\pi }}`$ |
| $`3(mod4)`$ | $`{\displaystyle \frac{L+\pi }{L\pi }}`$ | $`{\displaystyle \frac{\left(1+i\right)\pi +\left(1+i\right)L}{\left(1+i\right)\pi +\left(1i\right)L}}`$ | $`{\displaystyle \frac{L+\pi }{\left(1+i\right)\pi +\left(1i\right)L}}`$ |
| $`4(mod4)`$ | $`{\displaystyle \frac{L}{\pi }}`$ | $`{\displaystyle \frac{\pi iL}{\pi +iL}}`$ | $`{\displaystyle \frac{L}{\pi +iL}}`$ |
Note. The shorthand $`L=\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)`$ is used in this table.
The above analysis refers to $`l=0`$. For all other values $`l>0`$, the coupling will be weak, so that the phase shifts will be given by the values of Eq. (63), with the condition that $`\lambda =\lambda _{_{(\mathrm{gs})}}^{()}`$; then,
$$\delta _l^{(D_0)}|_{l0}=\left[\left(l+\nu _0\right)\sqrt{l\left(l+2\nu _0\right)}\right]\frac{\pi }{2},$$
(174)
and
$$S_l^{(D_0)}(k)|_{l0}=\left(1\right)^l\mathrm{exp}\left(i\nu _0\pi \right)\mathrm{exp}\left[i\pi \sqrt{l\left(l+2\nu _0\right)}\right],$$
(175)
which are scale-invariant expressions.
A number of consequences follow from Eqs. (171) and (172):
1. The unique pole of the scattering matrix (171) corresponds to the unique bound state.
2. The phase shifts can be renormalized using a floating renormalization scale $`\mu `$, as an alternative to ground-state renormalization. Then the $`s`$-wave phase shift reads simply
$$\frac{1}{\mathrm{tan}\mathbf{(}\text{ }\delta _0^{(D_0)}(k)\pi \nu _0/2\mathbf{)}\text{ }}=\frac{1}{\mathrm{tan}\mathbf{(}\text{ }\delta _0^{(D_0)}(\mu )\pi \nu _0/2\mathbf{)}\text{ }}+\frac{1}{\pi }\mathrm{ln}\left(\frac{k}{\mu }\right)^2.$$
(176)
3. Equations (173), (174), (B8), and (B9) give the differential scattering cross section
$$f_k^{(D_0)}(\mathrm{cos}\theta )=\frac{N_{\nu _0}}{2ik^{(D_01)/2}}\frac{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)\left(e^{i\nu _0\pi }1\right)+i\pi \left(e^{i\nu _0\pi }+1\right)}{\mathrm{ln}\left(k^2/|E_{_{(\mathrm{gs})}}|\right)i\pi }+\underset{k}{\overset{(D_0)}{\stackrel{}{f}}}(\mathrm{cos}\theta ),$$
(177)
where
$`\underset{k}{\overset{(D_0)}{\stackrel{}{f}}}(\mathrm{cos}\theta )={\displaystyle \frac{N_{\nu _0}}{2ik^{(D_01)/2}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}`$ $`\left({\displaystyle \frac{l}{\nu _0}}+1\right)`$ (178)
$`\times `$ $`\left\{\left(1\right)^le^{i\nu _0\pi }e^{i\pi \sqrt{l\left(l+2\nu _0\right)}}1\right\}C_l^{(\nu _0)}(\mathrm{cos}\theta ).`$ (179)
4. All the relevant quantities are logarithmic with respect to the energy and agree with the predictions of generalized dimensional analysis .
5. Equations (171), (172), and (173) relate the bound-state and scattering sectors of the theory and show that the inverse square potential is renormalizable.
## V CONCLUSIONS
In this paper, we have uncovered a number of remarkable analogies between the two-dimensional delta-function and inverse square potentials. In addition to displaying their characteristic transmuting behavior, with all the ensuing implications, we have explicitly seen that they share the following properties:
(i) Unusual boundary conditions at the origin.
(ii) Characteristic critical couplings (coincident for $`D_0=2`$) that determine the possible regimes of each potential.
(iii) Only one bound state in the renormalized theory, even though this is achieved through different mechanisms (the delta-function potential always generates a unique state, while the inverse square potential annihilates all the regularized excited states by exponential suppression).
(iv) Almost identical ground-state wave functions (up to the normalization constant).
(v) Similar $`s`$-wave scattering matrix elementsโthey are proportional, differing only upon an extra $`D_0`$-dependent phase factor for the inverse square potential.
(vi) Characteristic logarithmic behavior of $`s`$-wave scattering quantities.
It should be noticed that the bound-state sectors look essentially identical in both theories, while the $`s`$-wave scattering sectors are identical for $`D=2`$ and almost identical for $`D2`$.
However, there a number of differences as well. Due to its zero-range nature, the two-dimensional delta function only scatters $`s`$ waves, while the inverse square potential, due to its infinite range, scatters all other angular-momentum channels in a scale-invariant (energy-independent) way.
Moreover, these results are independent of the regularization technique; in particular, they are in perfect agreement with the $`D_0`$-dimensional generalization of the cutoff-renormalization method of Ref. .
Finally, the techniques used in this paper could be easily generalized. For example, one could consider the generalized inverse square potential
$$V(๐ซ)=\lambda \frac{v(\mathrm{\Omega }^{(D)})}{r^2},$$
(180)
with a dimensionless function $`v(\mathrm{\Omega }^{(D)})`$ that depends on the $`D`$-dimensional solid angle $`\mathrm{\Omega }^{(D)}`$; the angular part of the solution could be properly modified, but the basic scaling relationships would remain the same. In principle, this strategy could be conveniently used for the dipole potential and for other forms of angular dependence in Eq. (180).
## A ROTATIONAL INVARIANCE: CENTRAL POTENTIALS IN $`D`$ DIMENSIONS
In this appendix we will enforce the condition of rotational invariance and use the notation and definitions of hyperspherical coordinates as introduced in Ref. .
As usual, a central potential $`V(r)`$ is defined to be rotationally invariant; thus, its functional form is independent of the angular variables in $`D`$-dimensional hyperspherical coordinates. Its associated symmetry group is $`SO(D)`$, with a quantum mechanical representation generated by the $`D(D1)/2`$ generalized angular momentum operators $`x_ip_kx_kp_i`$ (with $`i<k`$). The quadratic Casimir operators
$`L_j^2`$ $`=`$ $`{\displaystyle \underset{k>i=j}{\overset{D1}{}}}\left(x_ip_kx_kp_i\right)^2`$ (A1)
$`=`$ $`\mathrm{}^2{\displaystyle \underset{k=j}{\overset{D1}{}}}\left({\displaystyle \underset{i=j}{\overset{k1}{}}}\mathrm{sin}^2\theta _i\right)^1\left[{\displaystyle \frac{^2}{\theta _k^2}}+(Dk1)\mathrm{cot}\theta _k{\displaystyle \frac{}{\theta _k}}\right]`$ (A2)
(with $`j=1,\mathrm{}D1`$) commute with all the generators of the Lie algebra and with all possible rotations. Each $`L_j^2`$ represents the โtotalโ $`D`$-dimensional angular momentum squared in the subspace spanned by the Cartesian coordinates $`(x_j,\mathrm{},x_D)`$, and is characterized by the properties that it commutes with the Hamiltonian of any central potential and that it satisfies the eigenvalue equation
$$L_j^2|E,L>=l_j(l_j+Dj1)\mathrm{}^2|E,L>,$$
(A3)
where the eigenstates $`|E,L>`$ are labeled by the energy and by the collective index of generalized angular momentum quantum numbers
$$L=\left(ll_1,l_2,\mathrm{},l_{D2},ml_{D1}\right).$$
(A4)
The eigenstates $`|E,L>`$ lead to the generalization of the usual 3D factorization of the position wave function, which can be written as
$$\mathrm{\Psi }_{EL}(r,\mathrm{\Omega }^{(D)})=<r,\mathrm{\Omega }^{(D)}|E,L>=R_{El}(r)Y_L(\mathrm{\Omega }^{(D)}),$$
(A5)
in terms of the hyperspherical harmonics $`Y_L(\mathrm{\Omega }^{(D)})=<\mathrm{\Omega }^{(D)}|L>`$. Notice that the peculiar number $`m`$ is associated with rotations on the $`(x_{D1},x_D)`$ plane, which are characterized by the azimuthal angle $`\varphi \theta _{D1}`$; the corresponding operator is usually chosen to be
$$L_{D1}=x_{D1}p_Dx_Dp_{D1}=\frac{\mathrm{}}{i}\frac{}{\varphi },$$
(A6)
instead of $`L_{D1}^2`$, with eigenvalues $`m\mathrm{}`$ (both positive and negative). In addition, the generalized angular momentum quantum numbers satisfy the constraints
$$0|m|l_{D2}l_{D3}\mathrm{}l_3l_2l.$$
(A7)
Moreover, the angular part of the Laplacian ,
$$\mathrm{\Delta }_{\mathrm{\Omega }^{(D)}}=\underset{j=1}{\overset{D1}{}}\left[\left(\underset{k=1}{\overset{j1}{}}\mathrm{sin}^2\theta _k\right)\mathrm{sin}^{Dj1}\theta _j\right]^1\frac{}{\theta _j}\left(\mathrm{sin}^{Dj1}\theta _j\frac{}{\theta _j}\right),$$
(A8)
from Eq. (A2), can be simply written in terms of the total angular momentum $`L^2=L_1^2`$, namely,
$$\mathrm{\Delta }_{\mathrm{\Omega }_D}=\frac{L^2}{\mathrm{}^2},$$
(A9)
so that Eq. (A3) leads to
$$\mathrm{\Delta }_{\mathrm{\Omega }^{(D)}}Y_L(\mathrm{\Omega }^{(D)})=l(l+D2)Y_L(\mathrm{\Omega }^{(D)}).$$
(A10)
In particular, the solutions of the angular part of Laplaceโs equation that are independent of all angles but $`\theta _1`$ are the hyperspherical harmonics with $`l_2=\mathrm{}=l_{D2}=m=0`$, for which Eq. (A10) reduces to the ultraspherical differential equation, with regular solutions given by the Gegenbauer (or ultraspherical) polynomials $`C_l^{(\nu )}(\mathrm{cos}\theta _1)`$, defined in terms of its generating function (see Refs. )
$$\left(12tz+z^2\right)^\nu =\underset{l=0}{\overset{\mathrm{}}{}}C_l^{(\nu )}(t)z^l.$$
(A11)
This is all the background needed for our work; the reader may consult Ref. for general proofs and Refs. for explicit expressions of the hyperspherical harmonics.
Equation (A10) justifies a posteriori the use of the notation $`R_{El}(r)`$ introduced in Eq. (A5), which assumes that $`R(r)`$ depends only on the angular momentum quantum number $`l`$ (but not on $`l_2,\mathrm{},l_{D2},m`$), in addition to the energy $`E`$. In effect, after isolating the angular factor that is common to all central potentials, the wave function $`R_{El}(r)`$ satisfies the equation
$$\left[\mathrm{\Delta }_r^{(D)}\frac{l(l+D2)}{r^2}+V(r)\right]R_{El}(r)=ER_{El}(r),$$
(A12)
with a radial Laplacian
$`\mathrm{\Delta }_r^{(D)}`$ $`=`$ $`{\displaystyle \frac{1}{r^{D1}}}{\displaystyle \frac{}{r}}\left(r^{D1}{\displaystyle \frac{}{r}}\right)`$ (A13)
$`=`$ $`{\displaystyle \frac{1}{r^{(D1)/2}}}{\displaystyle \frac{d^2}{dr^2}}\left[r^{(D1)/2}\right]+{\displaystyle \frac{(D1)(D3)}{4r^2}}.`$ (A14)
As a result, the function
$$u_{El}(r)=R_{El}(r)r^{(D1)/2}$$
(A15)
satisfies a one-dimensional Schrรถdinger equation
$$\left[\frac{d^2}{dr^2}+V(r)+\frac{\mathrm{\Lambda }_{l,D}}{r^2}\right]u_{El}(r)=Eu_{El}(r),$$
(A16)
in which the centrifugal barrier is characterized by the effective coupling constant
$`\mathrm{\Lambda }_{l,D}`$ $`=`$ $`l(l+D2)+(D1)(D3)/4`$ (A17)
$`=`$ $`(l+\nu )^21/4,`$ (A18)
where $`\nu `$ is the variable defined in Eq. (8). In our work, Eq. (A18) is most useful, particularly because, for the potentials of interest in this work, it provides a straightforward connection with a family of Bessel differential equations.
Equation (A18) depends on the number of dimensions of the space only through the combination $`l+\nu `$; this amounts to the remarkable phenomenon known as interdimensional dependence : the solutions for any two problems related via $`l+\nu =l^{}+\nu ^{}`$, with $`\nu \nu ^{}`$ (i.e., $`DD^{}`$) are identical.
A remark about notation is in order. In most contexts, it is customary to drop the subscript $`E`$ labeling the wave functions in Eqs. (A5), (A12), (A15), and (A16), i.e., to write $`R_{El}(r)R_l(r)`$ and $`u_{El}(r)u_l(r)`$. Alternatively, for the bound-state sector, one often writes $`R_{n_rl}(r)`$, $`u_{n_rl}(r)`$, and $`E_{n_rl}`$, where $`n_r`$ is the radial quantum number.
Finally, in order to have a well-defined problem, boundary conditions are needed both at infinity and at the origin. At infinity, the bound-state solutions should go to zero in order to ensure their integrability, while the scattering solutions are subject to the usual requirements . On the other hand, the boundary condition at $`r=0`$ requires further analysis.
In fact, the boundary condition at the origin is the basis for the classification of potentials into the regular and singular families, as discussed in Appendix C. In this framework, regular and semi-regular potentials are characterized by the limit
$$r^2V(r)\stackrel{r0}{}0,$$
(A19)
so that, near the origin, both the potential and total energy terms are negligible and the limiting form of the wave function becomes asymptotically a solution of the radial part of Laplaceโs equation, i.e.,
$$R_l(r)\stackrel{(r0)}{}\frac{u_l(r)}{r^{(D1)/2}}=\mathbf{\{}\text{ }r^l\text{,}\text{ }r^{(l+2\nu )}\mathbf{\}}\text{ },$$
(A20)
where $`\mathbf{\{}`$ , $`\mathbf{\}}`$ stands for linear combination. In Eq. (A20), the first component is acceptable but the second one should be discarded because $`_D^2\left[Y_L(\mathrm{\Omega }_D)/r^{l+D2}\right]`$ is proportional to the multipole density of order $`l`$, i.e., it involves derivatives of the delta function $`\delta ^{(D)}(๐ซ)`$ of order $`l`$. Therefore, for regular potentials, there is a criterion for the selection between the two linearly independent solutions of Eq. (A16), and this provides the boundary condition
$$R_l(r)r^l.$$
(A21)
In practice, it is sufficient to consider the weaker boundary condition
$$u_l(0)=0.$$
(A22)
As discussed in Subsection III B, the source of the unusual properties of critical (dimensionally transmuting) and singular potentials is the โlossโ of the boundary condition (A22).
## B ROTATIONAL INVARIANCE: PARTIAL-WAVE ANALYSIS IN $`D`$ DIMENSIONS
For our work, we need another aspect of rotational invariance in $`D`$ dimensions: the expansion in $`D`$-dimensional partial waves, which is developed next. A spherical wave state $`|E,L>`$ is represented by a solution of the radial Schrรถdinger equation for a free particle; then Eq. (A12), when $`V(r)=0`$, has solutions proportional to the Bessel functions $`Z_{l+\nu }(kr)`$, with $`\nu =D/21`$, where $`Z=J`$, or $`N`$, or $`H^{(1,2)}`$; explicitly,
$$R_l(r)z_{l,\nu }(x)=\sqrt{\frac{\pi }{2}}x^\nu Z_{l+\nu }(x),$$
(B1)
where the ultraspherical Bessel functions admit the asymptotic expansions
$$z_{l,\nu }(x)\stackrel{(x\mathrm{})}{}x^{(D1)/2}\tau \left(x+\gamma _Dl\frac{\pi }{2}\frac{\pi }{2}\right),$$
(B2)
in which $`\tau (\xi )`$ is the corresponding trigonometric function (i.e., $`\mathrm{cos}\xi `$ for $`j(\xi )`$, $`\mathrm{sin}\xi `$ for $`n(\xi )`$, and $`\mathrm{exp}(\pm i\xi )`$ for $`h^{(1,2)}(\xi )`$), and $`\gamma _D=(3D)\pi /4`$ (from Ref. ). In particular, the regular free-particle solution of Eq. (A12) (based on its behavior at the origin) is provided by $`j_{l,\nu }(kr)`$. On the other hand, the angular part leads to the Gegenbauer (or ultraspherical) polynomials $`C_l^{(\nu )}(\mathrm{cos}\theta )`$ defined through Eq. (A11). In short, the regular solution is of the form $`j_{l,D/21}(kr)C_l^{(D/21)}(\mathrm{cos}\theta )`$.
The partial-wave expansion for central potentials in $`D`$ dimensions then proceeds in complete analogy to the three-dimensional case . This can be accomplished by considering the transition from a plane-wave state $`|๐ค>`$, represented by the wave function $`e^{ikr\mathrm{cos}\theta }`$ (we will choose $`\theta \theta _1`$ for the sake of simplicity), to a spherical-wave state $`|E,L>`$. The coefficients of the transition can be found from the identity (Ref. , p. 363)
$$e^{ix\mathrm{cos}\theta }=\mathrm{\Gamma }(\nu )\left(\frac{x}{2}\right)^\nu \underset{l=0}{\overset{\mathrm{}}{}}\left(l+\nu \right)i^lJ_{l+\nu }(x)C_l^{(\nu )}(\mathrm{cos}\theta ),$$
(B3)
with $`x=kr`$; this implies that
$$e^{ikr\mathrm{cos}\theta }=N_\nu \underset{l=0}{\overset{\mathrm{}}{}}\left(\frac{l}{\nu }+1\right)i^lj_{l,\nu }(kr)C_l^{(\nu )}(\mathrm{cos}\theta ),$$
(B4)
which is the $`D`$-dimensional generalization of Rayleighโs formula, with
$$N_\nu =2^\nu \mathrm{\Gamma }(\nu +1)\sqrt{\frac{2}{\pi }}.$$
(B5)
Equation (B4) straightforwardly reduces to the familiar result for $`D=3`$ ($`\nu =1/2`$). The case $`D=2`$ ($`\nu =0`$) appears to be singular, but also reproduces the known results when the following replacements are made: (i) $`C_l^{(0)}(t)=lim_{\nu 0}C_l^{(\nu )}(t)/\nu `$ (Ref. ); (ii) $`C_l^{(0)}(\mathrm{cos}\theta )=2\mathrm{cos}(l\theta )/l`$ for $`l0`$ (this is proportional to a Chebyshev polynomial, Ref. ) but $`C_0^{(0)}(\mathrm{cos}\theta )=1`$; and (iii) most often, the sum is extended from $`m=\mathrm{}`$ to $`m=\mathrm{}`$, with $`m=\pm l`$, and the factor of 2 in (ii) is removed (otherwise this factor is kept as the Neumann number, $`ฯต_l=2`$ for $`l0`$, but $`ฯต_0=1`$).
As the transition operator $`T`$ commutes with the generalized angular momentum operators, it has a diagonal form in the angular momentum eigenbasis, with elements $`T_l^{(D)}(k)`$; thus, one can expand its matrix elements $`<๐ค|T|๐ค^{}>`$ in terms of $`T_l^{(D)}(k)`$, with
$$T_l^{(D)}(k)=\frac{S_l^{(D)}(k)1}{2\pi i},$$
(B6)
where the scattering matrix elements
$$S_l^{(D)}(k)=\mathrm{exp}\left[2i\delta _l^{(D)}(k)\right]=\frac{1+i\mathrm{tan}\delta _l^{(D)}(k)}{1i\mathrm{tan}\delta _l^{(D)}(k)}$$
(B7)
are usually expressed in terms of the scattering phase shifts $`\delta _l^{(D)}(k)`$. Then the corresponding scattering amplitude $`f_k^{(D)}(\mathrm{cos}\theta )`$ admits a straightforward expansion
$$f_k^{(D)}(\mathrm{cos}\theta )=N_\nu \underset{l=0}{\overset{\mathrm{}}{}}\left(\frac{l}{\nu }+1\right)a_l^{(D)}(k)C_l^{(\nu )}(\mathrm{cos}\theta ),$$
(B8)
with an $`l`$th partial-wave amplitude
$$a_l^{(D)}(k)=\frac{\pi }{k^{(D1)/2}}T_l^{(D)}(k)=\frac{\mathrm{exp}\left[2i\delta _l^{(D)}(k)\right]1}{2ik^{(D1)/2}}$$
(B9)
that provides the asymptotic form of the wave function
$`\mathrm{\Psi }(๐ซ)\stackrel{(r\mathrm{})}{}{\displaystyle \frac{N_\nu }{(kr)^{(D1)/2}}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{l}{\nu }}+1\right)`$ $`\mathrm{exp}\left\{i\left[\delta _l^{(D)}(k)+l\pi /2\right]\right\}`$ (B10)
$`\times `$ $`\mathrm{sin}\left[kr+\gamma _Dl\pi /2+\delta _l^{(D)}(k)\right]C_l^{(\nu )}(\mathrm{cos}\theta ).`$ (B11)
As a practical matter, the scattering matrix can be computed directly from the asymptotic expansion of the exact solution to the problem, i.e.,
$$R_l(r)R_{El}(r)\stackrel{(r\mathrm{})}{}A_l^{(+)}h_{l,\nu }^{(1)}(kr)+A_l^{()}h_{l,\nu }^{(2)}(kr),$$
(B12)
which yields
$$S_l^{(D)}(k)=\frac{A_l^{(+)}}{A_l^{()}}.$$
(B13)
Finally, the total scattering cross section can be obtained directly by integration of $`|f_k^{(D)}(\mathrm{cos}\theta )|^2`$, with the result
$$\sigma _D(k)=\frac{2\mathrm{\Omega }_{D1}}{k^{D1}}\underset{l=0}{\overset{\mathrm{}}{}}\left(2l+2\nu \right)\frac{\mathrm{\Gamma }(l+2\nu )}{l!}\mathrm{sin}^2\delta _l^{(D)}(k).$$
(B14)
## C DUALITY TRANSFORMATION FOR POWER-LAW POTENTIALS
In this appendix, which we have adapted from Ref. , we will consider the class of central power-law potentials
$$V(r)=\mathrm{sgn}\left(\beta \right)\lambda r^\beta ,$$
(C1)
where the sign is chosen so that $`\lambda >0`$ corresponds to attractive potentials. According to their behavior at the origin, they can be classified into the following categories.
1. Regular potentials: $`\beta 0`$.
2. Semi-regular potentials: $`2<\beta <0`$.
3. Critically singular potential: $`\beta =2`$
4. Strictly singular potentials: $`\beta <2`$.
What physically characterizes the singular potentials is that the centrifugal barrier fails to be the dominant term at the origin; mathematically, the Hamiltonian loses its self-adjoint character . In particular, the critical potential is the one that generates dimensional transmutation.
The central theme of this appendix is the existence of a remarkable duality transformation $`๐`$, which relates the potentials in the regular and semi-regular families; in particular, $`๐`$ establishes a one-to-one correspondence between the exponents $`\beta `$ in the intervals $`(2,0]`$ and $`[0,\mathrm{})`$. As we will see below, if
$$๐\mathbf{(}\text{ }\mathrm{sgn}\left(\beta \right)\lambda r^\beta \mathbf{)}\text{ }=\mathrm{sgn}\left(\stackrel{~}{\beta }\right)\stackrel{~}{\lambda }r^{\stackrel{~}{\beta }},$$
(C2)
then the corresponding relation between exponents reads
$$(\beta +2)(\stackrel{~}{\beta }+2)=4.$$
(C3)
Equations (C2) and (C3) exhibit the following properties: (i ) $`๐`$ is idempotent; (ii) $`2<\beta <0`$ if and only if $`0<\stackrel{~}{\beta }<\mathrm{}`$, so that $`\mathrm{sgn}\left(\beta \right)=\mathrm{sgn}(\stackrel{~}{\beta })`$; (iii) the Coulomb potential ($`\beta =1`$) is the dual of the harmonic oscillator ($`\beta =2`$); (iv) its limiting exponents ($`\beta =2`$ and $`\beta =\mathrm{}`$) define the inverse square potential as the dual of the infinite hyperspherical potential; and (v) the constant potential or free-particle case is self-dual.
Let us now show that Eq. (C3) represents the only nontrivial duality transformation implemented by a scale transformation for arbitrary power-law potentials. In general, under the transformation
$$\{\begin{array}{c}r=f(\varrho )\hfill \\ u(r)=\stackrel{~}{u}(\varrho )g(\varrho )\hfill \end{array},$$
(C4)
from a radial wave function $`u(r)`$ to $`\stackrel{~}{u}(\varrho )`$, the Schrรถdinger equation
$$\left[\frac{d^2}{dr^2}+EV(r)\frac{(l+\nu )^21/4}{r^2}\right]u_l(r)=0$$
(C5)
is mapped into a transformed equation of the same form, without first-order derivative term, if and only if
$$\left[g(\varrho )\right]^2f^{}(\varrho ),$$
(C6)
where the prime, as usual, denotes the derivative; then
$$\left\{\frac{d^2}{d\varrho ^2}+\left[f^{}(\varrho )\right]^2\left[EV\mathbf{(}\text{ }f(\varrho )\mathbf{)}\text{ }\right]_1^2(\varrho )(l+\nu )^2+_2^{}(\varrho )_2^2(\varrho )\right\}\stackrel{~}{u}(\varrho )=0,$$
(C7)
where
$$_j(\varrho )=\frac{1}{j!}\left(\frac{d\mathrm{ln}}{d\varrho }\right)^jf(\varrho )$$
(C8)
(for $`j=1,2`$). Even though Eqs. (C4) and (C7) are fairly general and have many applications, the scale transformation
$$\{\begin{array}{c}r=\varrho ^\alpha \hfill \\ u(r)=\stackrel{~}{u}(\varrho )\varrho ^{(\alpha 1)/2}\hfill \end{array},$$
(C9)
leading to
$$\left\{\frac{d^2}{d\varrho ^2}+\alpha ^2\varrho ^{2(\alpha 1)}\left[EV(\varrho ^\alpha )\right]\frac{\alpha ^2(l+\nu )^21/4}{\varrho ^2}\right\}\stackrel{~}{u}(\varrho )=0,$$
(C10)
can immediately provide the desired connection for the power-law potentials of Eq. (C1). In effect, unless the trivial transformation $`\alpha =1`$ is allowed, the only way of bringing the middle term in Eq. (C10) to the required transformed form, $`\stackrel{~}{E}\stackrel{~}{V}(\varrho )`$, is to perform the exchange
$`V(r)`$ $`=`$ $`\mathrm{sgn}\left(\beta \right)\lambda r^\beta \stackrel{~}{E}`$ (C11)
$`E`$ $``$ $`\stackrel{~}{V}(\varrho )=\mathrm{sgn}\left(\stackrel{~}{\beta }\right)\stackrel{~}{\lambda }\varrho ^{\stackrel{~}{\beta }};`$ (C12)
this โcrossingโ implies that the exponents be constrained by
$$\alpha =\frac{\stackrel{~}{\beta }}{\beta }=\frac{2}{\beta +2},$$
(C13)
which is equivalent to the anticipated duality relation (C3). In conclusion, the required coordinate and wave function substitutions are
$$\{\begin{array}{c}r=\varrho ^{2/(\beta +2)}\hfill \\ u(r)=\stackrel{~}{u}(\varrho )\varrho ^{\beta /2(\beta +2)}\hfill \end{array},$$
(C14)
which transform Eq. (C5) into
$$\left[\frac{d^2}{d\varrho ^2}+\stackrel{~}{E}\mathrm{sgn}\left(\stackrel{~}{\beta }\right)\stackrel{~}{\lambda }\varrho ^{\stackrel{~}{\beta }}\frac{\left(\stackrel{~}{l}+\stackrel{~}{\nu }\right)^21/4}{\varrho ^2}\right]\stackrel{~}{u}_l(\varrho )=0,$$
(C15)
with a dual energy eigenvalue
$$\stackrel{~}{E}=\mathrm{sgn}\left(\beta \right)\lambda \alpha ^2,$$
(C16)
dual coupling
$$\stackrel{~}{\lambda }=\mathrm{sgn}\left(\beta \right)E\alpha ^2,$$
(C17)
and dual angular momentum quantum number $`\stackrel{~}{l}`$, such that
$$\stackrel{~}{l}+\stackrel{~}{\nu }=\alpha (l+\nu ),$$
(C18)
where $`\alpha `$ is explicitly given by Eq. (C13), and we also allow for the possibility of a change in the number of dimensions, according to $`\stackrel{~}{\nu }=\stackrel{~}{D}/21`$. In particular, Eq. (C15) shows that the dual potential is, simply,
$$\stackrel{~}{V}(\varrho )=E\alpha ^2\varrho ^{\stackrel{~}{\beta }}.$$
(C19)
Finally, the dimensionless form of Eq. (C15) can be obtained by introducing an inverse length scale from the original problem through the energy, i.e.,
$$\kappa =\sqrt{|E|};$$
(C20)
then the resulting transformation involves the dimensionless variables
$$\{\begin{array}{c}z=\kappa ^{1/\alpha }\varrho \hfill \\ w(z)=\kappa ^{(D+11/\alpha )/2}\stackrel{~}{u}(\kappa ^{1/\alpha }z)\hfill \end{array},$$
(C21)
in terms of which
$$\{\begin{array}{c}\kappa r=z^{2/(\beta +2)}\hfill \\ \kappa ^{D/2}u(r)=w(z)z^{\beta /2(\beta +2)}\hfill \end{array}$$
(C22)
and the transformed Schrรถdinger equation becomes
$$\left[\frac{d^2}{dz^2}\mathrm{sgn}\left(\beta \right)\lambda \alpha ^2\kappa ^{2/\alpha }+\sigma \alpha ^2z^{\stackrel{~}{\beta }}\frac{(\stackrel{~}{l}+\stackrel{~}{\nu })^21/4}{z^2}\right]w_l(z)=0,$$
(C23)
where $`\sigma =\mathrm{sgn}(E)`$. Two important remarks immediately follow from Eq. (C23): (i) the ensuing dimensionless energy equation,
$$\stackrel{~}{\eta }=\mathrm{sgn}\left(\beta \right)\lambda \alpha ^2|E|^{1/\alpha },$$
(C24)
is the basis for the relation between the energy eigenvalue problems of the corresponding dual potentials (for example, for $`\beta =1`$, $`\stackrel{~}{\beta }=2`$, $`\alpha =2`$, it provides the well-known connection between the Coulomb potential and the harmonic oscillator, in spaces of any number of dimensions); and (ii) the dimensionless dual potential
$$\stackrel{~}{๐ฑ}(z)=\sigma \alpha ^2z^{\stackrel{~}{\beta }}$$
(C25)
satisfies the sign relation $`\mathrm{sgn}(\stackrel{~}{๐ฑ})=\mathrm{sgn}(E)`$, which, combined with Eq. (C24), leads to a one-to-one correspondence between the bound-state sectors and between the scattering sectors of the dual potentials.
For our work, the transformation in the โneighborhoodโ of the inverse square potential amounts to $`\beta =(2ฯต)<0`$, with $`ฯต1`$, which implies that $`\stackrel{~}{\beta }=4/ฯต21`$ and $`\alpha =2/ฯต`$, thus converting Eq. (C23) into
$$\left[\frac{d^2}{dz^2}+\frac{4}{ฯต^2}\lambda \kappa ^ฯต+\sigma \frac{4}{ฯต^2}z^{4/ฯต2}\frac{(\stackrel{~}{l}+\stackrel{~}{\nu })^21/4}{z^2}\right]w_l(z)=0,$$
(C26)
where the potential energy term $`\sigma \mathrm{\hspace{0.17em}4}z^{4/ฯต2}/ฯต^2`$ is seen to behave as an infinite hyperspherical potential well in the limit $`ฯต=0^+`$, for $`E<0`$ (i.e., $`\sigma =1`$).
ACKNOWLEDGEMENTS
This research was supported in part by CONICET and ANPCyT, Argentina (L.N.E., H.F., and C.A.G.C.) and by the University of San Francisco Faculty Development Fund (H.E.C.). H.E.C. acknowledges the generous hospitality of the University of Houston while this article was written. |
warning/0003/math-ph0003010.html | ar5iv | text | # Configuration Spaces and the Topology of Curves in Projective Space
## Abstract
We survey and expand on the work of Segal, Milgram and the author on the topology of spaces of maps of positive genus curves into $`n`$-th complex projective space, $`n1`$ (in both the holomorphic and continuous categories). Both based and unbased maps are studied and in particular we compute the fundamental groups of the spaces in question. The relevant case when $`n=1`$ is given by a non-trivial extension which we fully determine.
ยง1 Introduction and Statement of Results
The topology of spaces of rational maps into various complex manifolds has been extensively studied in the past two decades. Initiated by work of Segal and Brockett in control theory, and then motivated by work of Donaldson in gauge theory, this study has uncovered some beautiful phenomena (cf. \[CM\], \[Hu\], \[L\]) and brought to light interesting relationships between various areas of mathematics and physics (cf. \[Hi\], \[BM\],\[BHMM\]).
Let $`C_g`$ (or simply $`C`$) denote a genus $`g`$ (compact) Riemann surface ($`๐^1`$ when $`g=0`$) and let $`V`$ be a complex projective variety. Both $`C`$ and $`V`$ come with the choice of preferred basepoints $`x_0`$ and $``$ respectively. The main focus of this paper is the study of the geometry of the space
$$\text{Hol}(C,V)=\{f:CV,f\text{holomorphic}\}$$
and more particularly its subspace of basepoint preserving maps
$$\text{Hol}^{}(C,V)=\{f:CV,f\text{holomorphic and}f(x_0)=\}$$
The space $`\text{Hol}^{}(C,V)`$ doesnโt depend up to homeomorphism on the choice of basepoint when $`V`$ has a transitive group of automorphisms for example. In that situation, the relationship between the based and unbased (or free) mapping spaces is given by the โevaluationโ
$$\text{Hol}^{}(C,V)\text{Hol}(C,V)\genfrac{}{}{0pt}{}{ev}{}V$$
which is a holomorphic fibration for $`V`$ a homogeneous space for example. Here $`ev`$ evaluates a map at the basepoint of $`C`$.
Remark 1.1: By choosing a Riemann surface, we fix the complex structure. We record the link with the theory of โpseudo-holomorphicโ curves of Gromov. There one studies the (differential) geometry of spaces of $`J`$-holomorphic curves into $`V`$, where $`V`$ is almost complex and $`J`$ is its almost complex structure. These $`J`$-holomorphic curves (of given genus $`g`$) correspond when $`J`$ is integrable to all holomorphic maps of $`C`$ (with varying complex structure) into $`V`$. It is now a theorem of Gromov that for a generic choice of $`J`$ on $`V`$, the space of all such curves (in a given homology class) is a smooth finite dimensional (real) manifold. Let $`(V,g)=\{(C_g,f)|f:CV\text{holomorphic}\}`$, then there is a projection of $`(V,g)`$, onto the moduli space of curves $`_g`$ sending $`(C,u)`$ to the isomorphism class of $`C`$. The spaces $`\text{Hol}(C,V)`$ appear then as โfibersโ of this projection.
We observe that $`\text{Hol}(C,V)`$ breaks down into connected components and we write $`\text{Hol}_A(C,V)`$ for the component of maps $`f`$ such that $`f_{}[C]=A`$, $`A`$ a given homology class in $`H_2(V)`$. When $`V`$ is simply connected, $`[C,V]=[S^2,V]=\pi _2(V)=H_2(V)=๐^r`$ for some rank $`r`$ and $`A`$ is completely determined by a multi-degree. The components $`\text{Hol}_A(C,V)`$ are generally (singular) quasi-projective varieties (see \[H2\] for example). When $`V=๐^n`$, we write $`\text{Hol}_A(C,๐^n)=\text{Hol}_k(C,๐^n)`$ for some $`k=\text{deg}A๐`$ and these are smooth manifolds as soon as $`k`$ is big enough (about twice the genus of $`C`$). Most of this paper is concerned with the study of these spaces.
Remark 1.2: In physics, spaces of maps between manifolds $`\text{Hol}(M,V)\text{Map}(M,V)`$ arise in connection with field theory or โsigma modelsโ. From the perspective of a physicist, a field on $`M`$ with values in $`V`$ is a map $`\varphi :MV`$. For example, in the case $`M=๐^3`$ and $`V=๐^3=๐^3\{\mathrm{}\}`$, a map $`\varphi :๐^3S^3`$ could be the field associated to some electrical charge in $`๐^3`$ (hence vector valued and extending to the point where the charge is located by mapping to $`\mathrm{}`$). Associated to a field there is an โenergyโ density or Lagrangian $``$ (eg. the harmonic measure $`(\varphi )=\frac{1}{2}d(\varphi )^2`$). To an energy density one can in turn associate an โactionโ which is defined as
$$S[\varphi ]=_M(\varphi )๐\mu (h)$$
where $`d\mu (h)`$ is the canonical volume measure associated to a metric $`h`$ on $`M`$. Physicists are usually interested in minimizing the action (to determine the dynamics of the system) and hence they are led to study the space of all extrema of this functional. It should be noted that in the case when $`V`$ is compact, Kahler, a well-known theorem of Eells and Wood asserts that the absolute minima of the energy functional on $`\text{Map}(C,V)`$ are the holomorphic maps (the critical points here being the harmonic maps).
It has been known since the work of Segal \[S\] that the topology of holomorphic maps of a given degree $`k๐`$ from $`C`$ to $`๐^n`$ compares well with the space of continous maps at least through a range increasing with $`k`$. He proved
Theorem 1.3: (Segal) The inclusion
$$\text{Hol}_k^{}(C,๐^n)\text{Map}_k^{}(C,๐^n)$$
induces homology isomorphisms up to dimension $`(k2g)(2n1)`$
A similar statement holds for unbased maps. To simplify notation, we write $`\text{Map}_k`$, $`\text{Hol}_k`$, $`\text{Map}_k^{}`$, $`\text{Hol}_k^{}`$ for the corresponding mapping spaces from $`C`$ into $`๐^n`$ .
When $`g=0`$ (the rational case) and $`n>1`$ the homology isomorphism in 1.3 can be upgraded to a homotopy equivalence (cf. \[CS\]). For $`g>0`$ however it is not known whether the equivalence in 1.3 holds in homotopy as well; i.e whether the pair $`(\text{Map}_k,\text{Hol}_k)`$ is actually $`(k2g)(2n1)`$ connected. This is strongly suspected to be true and in this note we give further evidence for this by showing that $`\text{Hol}_k^{}`$ and $`\text{Map}_k^{}`$ have isomorphic fundamental groups for all $`n`$ (the relevant case here is $`n=1`$). In fact we shall show
Theorem 1.4: Suppose $`k2g`$, we have isomorphisms
$$\pi _1(\text{Hol}_k^{}(C_g,๐^n))\pi _1(\text{Map}_0^{}(C_g,๐^n))\{\genfrac{}{}{0pt}{}{๐^{2g},\text{when}n>1}{G,\text{when}n=1}$$
where $`G`$ is a cyclic extension of $`๐^{2g}`$ by $`๐`$ generated by classes $`e_1,\mathrm{},e_{2g}`$ and $`\tau `$ such that the commutators
$$[e_i,e_{g+i}]=\tau ^2$$
and all other commutators are zero.
Remark<sup>1</sup><sup>1</sup>1J.D.S.Jones has recently informed the author that he has obtained a similar result a few years ago but which he didnโt publish.: Roughly speaking, the class $`\alpha `$ can be represented by the one parameter family of maps obtained by rotating roots around poles. Similarly the classes $`e_i`$ are obtained by rotating roots (or poles) around loops representing the homology generators of $`C`$.
Combining this result with a classical result of G. Whitehead (see ยง2), we deduce
Proposition 1.5: $`\pi _1(\text{Map}_d(C,S^2))`$ is generated by classes $`e_1,\mathrm{},e_{2g}`$ and $`\alpha `$ such that
$$\alpha ^{2|d|}=1,[e_i,e_{g+i}]=\alpha ^2$$
and all other commutators are zero. When $`d2g`$, $`\pi _1(\text{Hol}_d(C,๐^1))\pi _1(\text{Map}_d(C,S^2)).`$
This description for the continuous mapping space is an earlier result of Larmore-Thomas \[LT\]. As is clear from 1.5, the components of $`\text{Map}(C,๐^n)`$ for $`n=1`$ have different homotopy types. When $`n>1`$, the fundamental group is however not enough to distinguish between the components. A quick byproduct of our calculations however shows (ยง8)
Proposition 1.6: $`\text{Map}_k(C,๐^{2d})`$ and $`\text{Map}_l(C,๐^{2d})`$ have different homotopy types whenever $`l`$ and $`k`$ have different parity<sup>2</sup><sup>2</sup>2it is now verified that all (positive) components of $`\text{Map}(C,๐^n)`$ have different homotopy type; cf. \[KS\]..
The first few sections of this paper are written in a leisurely fashion and are meant in part to survey techniques and ideas in the field most of which carry the deep imprint of Jim Milgram (cf. ยง4).
In ยง2 we discuss the rational case (i.e. $`g=0`$) and give a short proof of a theorem of Havlicek on the homology of unbased self-morphisms of the sphere. In ยง3 we introduce a configuration space model (originally given in \[K2\]) for $`\text{Map}^{}(C,๐^n)`$ and use it to determine the homology of this mapping space. Let $`SP^k(M)=M^k/\mathrm{\Sigma }_k`$ where $`\mathrm{\Sigma }_k`$ is the cyclic group on $`k`$-letters acting by permutations (or equivalently the space of unordered $`k`$ points on $`M`$), and let $`SP_n^k(M)`$ be the subspace of $`k`$ points in $`C`$ no more than $`n`$ of which are the same ($`nk`$). When $`n=1`$, $`SP_1^k(M)=C_k(M)`$ is the standard configuration space of $`k`$ distinct points. When $`M`$ is a (connected) open manifold or with a (collared) boundary, these spaces can be stabilized (cf. ยง3) $`SP_n^k(M)SP_n^{k+1}(M)\mathrm{}`$ and we write $`SP_n^{\mathrm{}}(M)`$ for the (connected) direct limit. The following is discussed in ยง3
Theorem 1.7 \[K2\]: There is a map
$$SP_n^{\mathrm{}}(C)\genfrac{}{}{0pt}{}{S}{}\text{Map}_0^{}(C,๐^n)$$
which is a homotopy equivalence when $`n>1`$ and a homology equivalence when $`n=1`$.
This model for $`\text{Map}^{}`$ in terms of โconfigurations of bounded multiplicityโ turns out to relate quite well with the corresponding model for $`\text{Hol}^{}`$ and we indicate based on this another proof for Segalโs result (ยง5). Notice that 1.4 shows that the homology equivalence $`S`$ above cannot be upgraded to a homotopy equivalence when $`n=1`$ for already both spaces have different fundamental groups. Indeed the braid group $`\pi _1(SP_1^{\mathrm{}}(C_g))=\pi _1(C(C_g))`$ has a presentation quite different from the central extension given in 1.4.
Acknowledgements: We would like to take this occasion to deeply thank Jim Milgram for the many beautiful mathematics we have learned from him. We thank J. F. Barraud, Y. Hantout and J. Nagel for useful discussions we had. Special thanks to the PIms institute in Vancouver for its support while part of this work was carried out.
ยง2 Preliminaries: The Genus Zero Case
It is customary to write $`\text{Rat}(V)`$ for the basepoint preserving maps $`\text{Hol}^{}(S^2,V)`$. When $`V=๐^n`$, these spaces are now very well understood. Let
$$C_k(๐,S^{2n1})=\underset{0ik}{}F(๐,i)\times _{\mathrm{\Sigma }_i}(S^{2n1})^i/$$
be the standard labeled $`k`$-th filtration piece for the May-Milgram model of $`\mathrm{\Omega }^2S^{2n+1}\mathrm{\Omega }_0^2(๐^n)`$ (cf. \[C<sup>2</sup>M<sup>2</sup>\], \[K4\], etc). Here $`F(๐,i)`$ is the set of ordered $`i`$-tuples of disctinct points in $`๐`$ and $``$ is a standard basepoint identification identifying an $`i`$-tuple with labels to an $`i1`$ tuple by discarding a point an its label if the label is at basepoint. Let
$$SP^k(X)=\{k_ix_i,x_iX,k_i๐|x_ix_j\text{for}ij\text{and}k_i=k\}$$
be the $`k`$-th symmetric product of $`X`$ (see introduction). Let $`SP_n^k(X)`$ be the subset of $`SP^k(X)`$ obtained by restricting to $`k_in`$. We then have
Theorem 2.1 (\[CS\], \[K3\], \[GKY\]): For $`k1`$, there are maps and homotopy equivalences
$$\text{Rat}_k(๐^n)\genfrac{}{}{0pt}{}{}{}C_k(๐,S^{2n1})\genfrac{}{}{0pt}{}{}{}SP_n^{k(n+1)}(๐)$$
$`2.1(a)`$
whenever $`n>1`$. When $`n=1`$ the spaces are homologous only.
It is interesting to note that no map is known to induce such homology isomorphism between $`\text{Rat}_k(๐^n)`$ and $`SP_n^{k(n+1)}(๐)`$ ($`k>1`$). When $`n>1`$, the left-hand map in 2.1(a) is constructed explicitly in \[CS\] while the right-hand map is constructed in \[K3\]. Both configuration space models on either side of 2.1(a) are fairly amenable to calculations and from there the structure of $`\text{Rat}_k(๐^n)`$ can be made quite explicit (cf. \[BM\], \[C<sup>2</sup>M<sup>2</sup>\], \[K3\] and \[Ka\]).
The space of unbased rational maps is less well understood. The following is due to Havlicek (\[H1\])
Theorem 2.2: The Serre spectral sequence for the (holomorphic) fiber bundle
$$\text{Rat}_k(๐^1)\text{Hol}_k(๐^1)๐^1$$
$`2.2(a)`$
has the non-zero differential $`d_2(x)=2k\iota `$. The spectral sequence collapses with mod-$`p`$ coefficients whenever $`p=2`$ or $`p`$ divides $`k`$.
We give a short proof for the mod-2 collapse (a general proof and an extension of this result to $`\text{Rat}_k(๐^n)`$ can be found in \[KS\]). First of all, to see that 2.2(a) is indeed a fiber bundle, one simply observes that PSL$`{}_{2}{}^{}(๐)`$ acts transitively on $`๐^1`$ and if $`F`$ denotes the stabilizer of a point, then $`F`$ acts on $`\text{Rat}_k(๐^1)`$ (by postcomposition). One can see that $`\text{Hol}_k(๐^1)=\text{Rat}_k(๐^1)\times _F๐^1`$.
Towards the proof of 2.2 (and also in most of ยง7) we need the following classical result of G. Whitehead. Let $`X`$ be a based (connected) topological space (with basepoint $`x_0`$) and consider the evaluation fibration
$$\mathrm{\Omega }_f^nX_f^nX\genfrac{}{}{0pt}{}{ev}{}X$$
$`2.3`$
where $`ev(f)=f(x_0)`$, $`_f^nX=\text{Map}_f(S^n,X)`$ is the component of the total mapping space containing a given map $`f:S^nX`$ and $`\mathrm{\Omega }_f^nX`$ the subset of all maps $`g`$ such that $`g(x_0)=f(x_0)`$.
Theorem 2.4 \[W\]: The homotopy boundary in 2.3 $`:\pi _i(X)\pi _i(\mathrm{\Omega }_f^nX)\pi _{i+n}(X)`$ is given (up to sign) by the Whitehead product: $`\alpha =[\alpha ,f].`$
Let $`_k^2S^2=_f^2S^2`$ where $`f`$ is the standard degree $`k`$ map, and denote by $`aH^2(S^2)`$ the generator. Recall that $`\mathrm{\Omega }^2S^2๐\times \mathrm{\Omega }^2S^3`$ and so let $`e`$ be the generator in $`H^1(\mathrm{\Omega }_k^2S^2)H^1(\mathrm{\Omega }^2S^3)`$.
Lemma 2.5: The Serre spectral sequence for $`\mathrm{\Omega }_k^2S^2_k^2S^2S^2`$ has the (homology) differential $`d_2(a)=2ke`$. It collapses at $`E_2`$ with mod-2 coefficients.
Proof: For a fibration $`FE\genfrac{}{}{0pt}{}{p}{}S^2`$ we have the diagram
$$\begin{array}{cccc}\genfrac{}{}{0pt}{}{\pi _2(E,F)}{\pi _2(S^2)}& \genfrac{}{}{0pt}{}{}{}& \pi _1(F)& \\ h& & h& \\ H_2(E,F)& \genfrac{}{}{0pt}{}{}{}& H_1(F)& \\ p_{}& & & \\ H_2(S^2)& \genfrac{}{}{0pt}{}{\tau }{}& E^{0,1}& \end{array}$$
where $`E^{0,1}H_1(F)`$ and $`\tau `$ the transgression. Let $`k:S^2S^2`$ denote multiplication by $`k`$. From 2.4 and the diagram above, we deduce that
$$\tau :H_2(S^2)\pi _2(S^2)\genfrac{}{}{0pt}{}{k[\iota ,]}{}\pi _3(S^2)\genfrac{}{}{0pt}{}{ad}{}\pi _1(\mathrm{\Omega }_k^2S^2)\genfrac{}{}{0pt}{}{h}{}H_1(\mathrm{\Omega }_k^2S^2)=๐$$
$`\iota \pi _2(S^2)`$ is the generator. As is known for even spheres, $`h(ad([\iota ,\iota ]))=2`$ and this yields the differential.
The mod-2 collapse follows from the following very short argument (which we attribute to Fred Cohen): let $`E:S^2\mathrm{\Omega }S^3`$ be the adjoint map. Then there is a map of (horizontal) fibrations
$$\begin{array}{cccccc}\mathrm{\Omega }_k^2S^2& & _k^2S^2& \genfrac{}{}{0pt}{}{ev}{}& S^2& \\ \mathrm{\Omega }^2E& & & & E& \\ \mathrm{\Omega }_k^3S^3& & _k^2\mathrm{\Omega }S^3& & \mathrm{\Omega }S^3& \end{array}$$
The map $`E`$ is injective in integral homology while $`\mathrm{\Omega }^2E`$ is injective in mod-2 homology. So mod-2, both fiber and base inject in the bottom fibration which is trivial since $`\mathrm{\Omega }S^3`$ is (homotopy equivalent) to a topological group (and translation of basepoint gives the trivialization). The collapse follows in this case.
Proof of 2.2: We have inclusions $`\text{Rat}(๐^1)\text{Hol}(๐^1)\mathrm{\Omega }^2S^2`$, where $`\text{Rat}(๐^1)`$ is the subspace of all holomorphic maps sending the north pole $`\mathrm{}`$ in $`๐^1=๐\mathrm{}`$ to 1. An element of $`\text{Rat}(๐^1)`$, say of degree $`k`$, is given as a quotient $`\frac{p}{q}=\frac{z^k+a_{k1}z^{k1}+\mathrm{}+a_0}{z^k+b_{k1}z^{k1}+\mathrm{}+b_0}`$ where $`p`$ and $`q`$ have no roots in common. It is easy to see for example that $`\text{Hol}_1(๐^1)`$ corresponds to $`PSL(2,๐)`$, the automorphism group of $`๐^1`$ (which is up to homotopy $`RP^3`$), and that $`\text{Rat}_1(๐^1)=๐\times ๐^{}S^1`$. Consider now the map of fibrations
$$\begin{array}{ccc}\text{Rat}_k(๐^1)& & \mathrm{\Omega }_k^2S^2\\ & & \\ \text{Hol}_k(๐^1)& & _k^2S^2\\ & & \\ ๐^1& \genfrac{}{}{0pt}{}{=}{}& S^2\end{array}$$
$`2.6`$
According to Segal (theorem 1.3), the top inclusion is an isomorphism in homology up to dimension $`k`$. Moreover (cf. \[C<sup>2</sup>M<sup>2</sup>\], \[K1\]), $`H_{}(\text{Rat}_k(๐^1))`$ actually injects in $`H_{}(\mathrm{\Omega }_0^2S^2)`$ and it does so in the following nice way. Recall that $`\mathrm{\Omega }_0^2S^2\mathrm{\Omega }^2S^3=\mathrm{\Omega }^2\mathrm{\Sigma }^2S^1`$ and hence it stably splits as an infinite wedge $`_{j1}D_j`$ where the summands $`D_j`$ are given in terms of configuration spaces with labels. Stably one finds that
$$\text{Rat}_k(๐^1)_s\underset{j=1}{\overset{k}{}}D_j.$$
The map of fibers in 2.6 is then an injection and since the bases are same in that diagram, the theorem follows from 2.5 and a spectral sequence comparison argument.
ยง3 Configuration Space Models and the Higher Genus Case
In this section we describe two configuration space models for each of the mapping spaces $`\text{Hol}_k^{}(C,๐^n)`$ and $`\text{Map}_k^{}(C,๐^n)`$ for $`g1`$. A straightforward comparison between both models yields Segalโs stability result in ยง4.
First of all, a map $`f\text{Hol}_k(C,๐^n)`$ can be written locally in the form
$$z[p_0(z):\mathrm{}:p_n(z)]$$
where the $`p_i(z)`$ are polynomials of degree $`k`$ each and having no roots in common. The $`p_i`$โs are not global functions on $`C`$ (otherwise they would be constant) but rather local maps into $`๐`$ and hence sections of some line bundle. The roots of each $`p_i`$ give rise to an element $`D_iSP^k(C)`$ (so called positive divisor) and conversely $`D_i`$ only determines $`p_i`$ up to a non-zero constant (which will be determined if the maps are based). These $`D_i`$โs (the root data) cannot of course have a root in common and if we base our maps so that basepoint $``$ is sent to $`[1:\mathrm{}:1]`$ say, then none of the $`D_i`$โs contains the basepoint. Define
$$\text{Div}_k^{n+1}(X)=\{(D_0,\mathrm{},D_n)SP^k(X)^{n+1}|D_0\mathrm{}D_n=\mathrm{}\},$$
The previous discussion shows the existence of an inclusion
$$\varphi :\text{Hol}_k^{}(C,๐^n)\text{Div}_k^{n+1}(C)$$
which associates to a holomorphic map its root data. This correspondence has no inverse since it is not true in general that an (n+1) tuple of degree $`k`$-divisors with no roots in common gives rise to a (based) holomorphic map out of $`C`$. In fact, this can be understood as follows: the quotients $`\frac{p_i(z)}{p_j(z)}`$ are meromorphic maps on $`C`$ and so for $`ij`$, $`D_iD_j`$ must be the divisor of a function on $`C`$ (i.e. there is $`f:C๐`$ with roots at $`D_i`$ and poles at $`D_j`$). It is not surprising that this last condition is not satisfied for general pairs $`(D_i,D_j)`$. We say $`D_i`$ and $`D_j`$ are linearly equivalent if there is indeed a meromorphic function $`f`$ on $`C`$ such that $`(f):=`$zeroes of $`f`$-poles of $`f=D_iD_j`$. Note that when this is the case $`D_i`$ and $`D_j`$ have same degree.
Linear equivalence defines an equivalence relation and one denotes by $`J(C)`$ the set of all linearly equivalent divisors on $`C`$ of degree $`0`$. There is a map
$$\mu :SP^n(C)J(C),D[Dnx_0]$$
$`3.1`$
sending a divisor to the corresponding equivalence class. The above discussion then shows the existence of a homeomorphism
$$\left\{\genfrac{}{}{0pt}{}{\text{A (based) holomorphic}}{X๐^r\text{ of degree }k}\right\}\left\{\genfrac{}{}{0pt}{}{\genfrac{}{}{0pt}{}{\text{An }\left(n+1\right)\text{ tuple of positive divisors }}{D_i\text{ on }C|D_0\mathrm{}D_n=\mathrm{}}}{\text{deg}D_i=k\text{ and }\mu (D_i)=\mu (D_j)\text{ }}\right\}$$
To make this correspondence more precise, we need understand the map $`\mu `$ in 3.1. It turns out that: $``$ $`J(C)`$ is a $`g`$ dimensional complex torus (this is a non-trivial fact) and $`\mu `$ is a multiplicative map. In the case $`g=1`$, $`\mu `$ is the identity and if one identifies the curve with $`๐/L`$ for some lattice $`L`$, then the Abel-Jacobi condition $`\mu (z_i)=\mu (p_j),z_i,p_iC`$ is equivalent to $`z_i=p_j\text{Mod}(L)`$ ($`z`$ for zero and $`p`$ for pole). $``$ The preimage of a point $`[D]=\mu (D)J`$ is a projective space. To see this let $`(D)`$ be the set of all holomorphic maps on $`C`$ such that $`(f)+D0`$ (i.e. such that $`(f)+D_nSP^n(C)`$). This is a $`๐`$-vector space and since $`(\alpha f)=(f)`$, we have an identification and a map $`\mu ^1([D])=๐(D)SP^n(C),[f](f)+D`$. This turns out to be a holomorphic embedding (\[G\],ยง4, 4.3). $``$ The complex dimension of $`\mu ^1([D])`$ is denoted by $`r(D)`$ and a crucial aspect of the classical theory of algebraic curves is the computation of $`r(D)`$ as $`D`$ varies in $`SP^n(C)`$. A very useful interpretation of this dimension is as follows: $`r(D)r`$ if and only for any $`r`$ points of $`C`$, there is a divisor $`D^{}`$ with $`\mu (D^{})=\mu (D)`$ and passing through these $`r`$ points. $``$ If $`\text{deg}D=n`$, then generically
$$r(D)=max\{0,ng\}.$$
The exact dimension for every $`D`$ is determined by Riemann-Roch. This dimension may jump up for certain โspecialโ divisors. However whenever $`n`$ exceeds $`2g2`$, there are no more jumps and $`r`$ is uniformally given by $`ng`$.
It turns out that for $`k`$ small (i.e. $`k<g`$), $`\text{Hol}_k^{}(C,๐^n)`$ is very dependent on the holomorphic structure put on the curve (it can naturally be empty). For example if $`C`$ is hyperelliptic, then
$$\text{Hol}_{2n+1}(C,๐^1)=\mathrm{},2n+1g\text{(see for instance [FK], Chap3)}$$
Also $`\text{Hol}_1(C,๐^n)`$ is always empty for all positive genus curves. In the range $`k2g1`$, there is however much better behavior (see \[KM\])
Lemma 3.2: Assume $`k2g1`$. Then $`\text{Hol}_k^{}(C,๐^n)`$ is a $`k(n+1)ng`$ complex manifold.
Remark 3.3: $`\text{Hol}_k^{}(C,๐^n)`$ has additionally the structure of a quasiprojective variety. Generally, components of $`\text{Hol}(C,V)`$ do not have a smooth structure (when $`V`$ is a smooth projective variety). This already fails for $`V=G(n,n+k)`$ (the Grassmaniann of $`n`$ planes in $`๐^{n+k}`$) when $`n>1`$ (see \[Ki\]). The โexpectedโ dimension of $`\text{Hol}_A(C,V)`$ (where $`A`$ is a fixed homology class in $`H_2(V)`$ and $`f_{}[C]=A`$ for all $`f\text{Hol}_A`$ can however be computed and at generic smooth points it is given by $`c_1.A+n(1g)`$. In our case $`c_1=n+1`$ ($`V=๐^n`$) and $`A`$ is $`k`$ fold the generator in $`H_2(V)`$.
ยง3.1 The configuration space model for $`\text{Map}^{}(C,๐^n)`$
In 1.7 we pointed out to the existence of a model for $`\text{Map}(C,๐^n)`$ in terms of bounded multiplicity symmetric products $`SP_n^k()`$ defined in ยง2. More precisely, consider $`CU`$ where $`U`$ is a little open disc around the basepoint $``$ ($`C`$ isotopy retracts onto $`CU`$). Let $`U_k`$ be a nested sequence of neighborhood retracting onto $``$, then one defines the space $`SP_n^{\mathrm{}}(C)`$ as the direct limit of inclusions $`SP_n^k(CU_k)SP_n^{k+1}(CU_{k+1})`$, $`DD+x_k,x_kU_kU_{k+1}`$ (here $`\{x_k\}`$ is a sequence of distinct points converging to $``$).
One can map $`SP_n^{\mathrm{}}(C)`$ to the based mapping space by scanning
$$S:SP_n^{\mathrm{}}(C)\text{Map}_0^{}(C,๐^n)$$
$`3.4`$
This map is given as follows (details in \[K2\]): identify canonically each neighborhood $`D(x)`$ of $`xC`$ with a closed disc $`D^2`$ (this is possible because $`C`$ is parallelizable). It follows that to every configuration in $`SP_n^{\mathrm{}}(C)`$ one can restrict to $`D(x)`$ and see a configuration in $`SP_n^{\mathrm{}}(D(x),D(x))=SP_n^{\mathrm{}}(D^2,D^2)`$ (the relative construction means that configurations are discarded when they get to the boundary). It can be checked that $`SP_n^{\mathrm{}}(D^2,D^2)=SP^n(S^2)=๐^n`$ and hence 3.4. It turns out that the map $`S`$ in 3.4. is a homotopy equivalence when $`n>1`$ or a homology equivalence when $`n=1`$ (cf. \[K2\]).
Lemma 3.5: Assume $`g1`$, $`kn1`$ and $`C^{}=C`$. Then the collar inclusion $`SP_n^k(C^{})SP^{\mathrm{}}(C^{})`$ induces an epimorphism on $`H_1`$ (for $`n>1`$ it is in fact an isomorphism).
Proof: A cohomology class in $`H^i(C^{},๐)`$ is represented by (the homotopy class) of a map $`C^{}K(๐,i)`$ and since the target is an abelian topological group, this map extends multiplicatively to
$$C^{}\genfrac{}{}{0pt}{}{i_1}{}SP_n^k(C^{})\genfrac{}{}{0pt}{}{i_2}{}SP^k(C^{})K(๐,i)$$
and hence gives rise to a class in $`H^i(SP_n^k(C^{}))`$. The โinclusionโ $`i_1:C^{}SP_n^k(C^{})`$ (constructed earlier) is then surjective in cohomology and hence injective in homology. Since $`\pi _1(SP^k(C^{}))=H_1(SP^k(C^{}),๐)=H_1(C^{};๐)`$, the composite $`i_2i_1`$ must be an isomorphism on $`H_1`$. This then implies that $`H_1(SP_n^k(C^{}))H_1(SP^k(C^{}))`$ is necessarily surjective. It can be checked that as soon as $`n2`$, $`\pi _1(SP_n^k(C^{}))`$ abelianizes (\[K3\]) and from there that $`H_1(SP_n^k(C^{}))H_1(SP^k(C^{}))=๐^{2g}`$.
Lemma 3.6: Assume $`g1`$. The map $`\text{Map}_k^{}(C,๐^n)\genfrac{}{}{0pt}{}{\alpha }{}\text{Map}_k^{}(C,๐^{\mathrm{}})(S^1)^{2g}`$, induced from post-composition with the inclusion $`๐^n๐^{\mathrm{}}`$, is an isomorphism at the level of $`H_1`$ when $`n>1`$ and a surjection when $`n=1`$.
Proof: Consider the following homotopy diagram
$$\begin{array}{ccc}SP_n^{\mathrm{}}(C)& \genfrac{}{}{0pt}{}{}{}& SP^{\mathrm{}}(C)\\ & & \\ \text{Map}_0^{}(C,๐^n)& \genfrac{}{}{0pt}{}{\alpha }{}& \text{Map}_0^{}(C,๐^{\mathrm{}})\end{array}$$
The top map is an isomorphism on $`H_1`$ (and only a surjection when $`n=1`$) according to the previous lemma. Since both vertical maps are isomorphisms on $`H_1`$ as well (by 3.4), the claim follows.
ยง4 A Spectral Sequence of Milgram
The following spectral sequence appears in special cases in \[BCM\], \[B\], \[K1\] and \[KM\] and the main ideas trace back to \[M1\]. Let $`X`$ be a space with basepoint $``$ and let $`E(X)SP^{\mathrm{}}(^kX)`$ be a submonoid of $`SP^{\mathrm{}}(^kX)=_kSP^{\mathrm{}}(X)`$. Given a map $`XE(X)`$, it can always be extended to a map $`\nu _0:SP^r(X)E(X)`$ (additively) and then to a map
$$\nu :\underset{r1}{}SP^r(X)\times EE,\nu (x,y)=\nu _0(x)+y$$
Our interest is to study the complement $`Par(X)=E(X)\text{Im}(\nu )`$ (or Particle space). We reserve the notation $`C(X)`$ for the standard configuration space of distinct points (see 4.1 below). Such generalized families of configuration spaces are studied in \[K2\].
Examples 4.1: $``$ (i) Let $`k=1`$ and consider the map $`ME(M)=SP^{\mathrm{}}(M)`$, $`x2x`$. Points in Im($`\nu `$) are finite sums $`n_ix_i`$ where $`n_i2`$ for at least one index $`i`$. In this case
$$Par(M)=C(M)=\{x_i,x_ix_j\text{for}ij\}$$
Similarly we can map $`ME`$ by sending $`x(n+1)x`$ and in this case the discriminant space is the space $`SP_n^{\mathrm{}}(M)`$ described in 3.4. $``$ (ii) The divisor space $`\text{Div}^n(M)`$ introduced in ยง3 is the particle space obtained as the complement of the diagonal map
$$M\stackrel{n}{}SP^{\mathrm{}}(M)=E(M),x(x,x,\mathrm{},x)$$
$``$ (iii) If $`\mu :MG`$ is a map into an abelian group $`G`$, then $`\mu `$ can be extended to $`SP^{\mathrm{}}(M)`$ and one can define
$$E(M,\mu )=\{(D_1,\mathrm{},D_k)\underset{k}{}SP^{\mathrm{}}(M)|\mu (D_i)=\mu (D_j)\text{for all}i,j\}$$
This is a submonoid since by construction $`\mu (D+D^{})=\mu (D)+\mu (D^{})`$. The main example we study in this paper is when $`M=C`$ is a curve and $`\mu `$ is its Abel-Jacobi map.
Consider the following construction
$$DE(M)=E\times _\nu SP^{\mathrm{}}(cM)$$
where $`cM`$ is the cone on $`M`$ and where the twisted product $`\times _\nu `$ means the identification via $`\nu `$ of the points
$$(\stackrel{}{\zeta },(t_1,z_1),\mathrm{},(t_l,z_l))(\stackrel{}{\zeta }+\nu (z_i),(t_1,z_1)\mathrm{}\widehat{(t_i,z_i)}\mathrm{}(t_l,z_l))$$
whenever $`t_i=0`$ is at the base of the cone (here hat means deletion). The identification above โcones offโ the image of $`\nu `$ in $`E`$ and so we expect that $`H_{}(DE)=H_{}(E/Im\nu )`$. Write
$$E_{n_1,\mathrm{},n_k}=E(M)\left(SP^{n_1}(M)\times \mathrm{}\times SP^{n_k}(M)\right)$$
and suppose $`\nu :ME(M)`$ lands in $`E_{l_1,\mathrm{},l_k},l_j1`$. We can then consider the filtration of $`DE`$ by subspaces
$$DE_{k_0,k_1,\mathrm{},k_n}(C_g)=\underset{\genfrac{}{}{0pt}{}{i_j+ll_jk_j}{1jk}}{}E_{i_0,i_1,\mathrm{},i_n}\times _\nu SP^l(cC_g)$$
There are (well-defined) projection maps
$$p_{k_0,k_1,\mathrm{},k_n}:DE_{k_0,k_1,\mathrm{},k_n}E_{k_0,k_1,\mathrm{},k_n}/\left\{\text{Image}(\nu )\right\}$$
sending $`(v_1,\mathrm{},v_s,(t_1,w_1),\mathrm{},(t_r,w_r))`$ to $`(v_1,\mathrm{},v_s)`$. These maps are acyclic and hence induce isomorphisms in homology (the proof in the case of the standard configuration spaces 4.1 (i) is given in \[BCM\], lemma 4.6 and relies mainly on the fact that $`SP^l(cX)`$ is contractible for every $`l`$).
Letโs assume $`l_j=d,1jk`$ and write $`(D)E_{i,\mathrm{},i}=(D)E_i`$. Suppose $`Par_i(M)=E_iIm(\nu )E_i`$ is an oriented manifold of dimension $`N(i)`$ (it will be for all cases we consider here otherwise we can use $`๐_2`$ coefficients). Then by Alexander-Poincarรฉ duality we have an isomorphism $`\stackrel{~}{H}^{N(i)}(Par_i(M);๐)H_{}(E_i/Im(\nu );๐)`$ for commutative rings $`๐`$. Combining this with the previous paragraph we get
Proposition 4.2: There are isomorphisms
$`\stackrel{~}{H}^{N(i)}(Par_i(M);๐)`$ $``$ $`H_{}(DE_i;๐)`$
$`\stackrel{~}{H}^{N(i)}(Par_i(M);๐)`$ $``$ $`H_{}(DE_i/{\displaystyle \underset{i}{}}DE_{i,\mathrm{},i1,\mathrm{},i};๐)`$
Let $`LE_i`$ be the quotient $`E_{i,\mathrm{},i}/E_{i,\mathrm{},i1,\mathrm{},i},i1`$. The map $`\nu `$ is made out of maps
$$\nu :SP^r(M)\times E_iE_{i+rd}$$
and so we can consider the quotient model
$$QE_k=DE_{k,\mathrm{},k}/DE_{k,\mathrm{},k1,\mathrm{},k}=\underset{i+dj=k}{}LE_i\times _\nu (SP^j(cM)/SP^{j1}(cM))$$
One can filter this complex by $`j`$ (in which case one obtains an analog of the Eilenberg-Moore spectral sequence with $`E^2`$ term a Tor term; see \[K1\], \[BCM\]) or one can filter according to $`i`$ by letting
$$_r=\underset{\genfrac{}{}{0pt}{}{ir}{i+dj=r}}{}LE_i\times _\nu (SP^j(cM)/SP^{j1}(cM))$$
In this case we get the spectral sequence
Theorem 4.3: There is a spectral sequence converging to $`H^{N(k)}Par_k(M)`$ with $`E^1`$ term
$$E^1=\underset{\genfrac{}{}{0pt}{}{i+dj=k}{i1}}{}H_{}(LE_i;๐)H_{}(SP^j(\mathrm{\Sigma }M),SP^{j1}(\mathrm{\Sigma }M);๐)$$
and where the $`d^r`$ differentials are obtained from a chain approximation of the identification maps $`\nu :SP^j(M)\times LE_iLE_{i+d}`$, $`(x,y)\nu (x)+y`$.
Sketch of proof: (see \[K1\] and \[KM\] for details) Consider $`_r`$ as described above. Then a chain complex for $`_r`$ is given as
$$C_{}(LE_i)_\nu C_{}(SP^l(cM)/SP^{l1}(cM))$$
Here the symmetric product pairing as well as the identification given by $`t`$ can be chosen to be cellular. We need determine the homology of this complex. An interesting fact is that one can construct chain complexes
$$C_{}(SP^{\mathrm{}}(cM))=C_{}(SP^{\mathrm{}}(M))C_{}(SP^{\mathrm{}}(\mathrm{\Sigma }M))$$
$`4.4`$
where the identification above is given as a bigraded differential algebra isomorphism. This follows from the fact (cf.\[M1\])) that $`C_{}(SP^{\mathrm{}}(cM))`$ can be identified with the acyclic bar construction on $`C_{}(SP^{\mathrm{}}(M)`$ and that $`C_{}(SP^{\mathrm{}}(\mathrm{\Sigma }M))`$ can be identified with the reduced bar construction on $`C_{}(SP^{\mathrm{}}(M)`$ . Therefore the boundary on cells of $`_r`$ can be made explicit. First a cell in $`C_{}(SP^l(cM)/SP^{l1}(cM))`$ can be written as
$$c_{}|a_1|\mathrm{}|a_l|$$
(according to the decomposition 4.4) and the boundary decomposes into
$$=c_{}|a_1|\mathrm{}|a_l|+\nu _{}(c_{}a_1)|a_2|\mathrm{}|a_l|+c_{}_B(|a_1|\mathrm{}|a_l|)$$
$`4.5`$
where $`_B`$ is the usual bar differential. The induced boundary on $`_r/_{r+1}`$ (which describes $`d_0`$) is given by the first and last term and the homology of the complex $`:C_{}(_r/_{r+1})C_{}(_{r+1}/_{r+2})`$ is given by the expression in 4.3. Remains to identify the $`d^r`$ differentials and these are deduced by the middle term in 4.5 according to the filtration term in which they land.
ยง4.1 An application: Configurations with bounded multiplicity
The homology of $`SP_n^{\mathrm{}}(C^{}),C^{}=C`$ can be calculated from the filtration pieces $`SP_n^k(C)`$ as follows. Start by observing that the configuration space $`SP_n^{\mathrm{}}(C)`$ is the discrimant set in $`SP^{\mathrm{}}(C)`$ of the image of $`\nu :SP^{\mathrm{}}(C)SP^{\mathrm{}}(C)`$ which is multiplication by $`n+1`$ (as described in 4.1 (i)). In this particular case 4.3 takes the form (with field coefficients)
(4.6 ) There is a spectral sequence converging to $`H_{2k}(SP_n^k(C);๐
)`$, $`n1`$, with $`E_{i,j}^1`$ term (to which we refer as $`E^1(k)`$ )
$$\underset{\genfrac{}{}{0pt}{}{i+(n+1)j=k}{r+s=}}{}H_r(SP^i(C),SP^{i1}(C);๐
)H_s(SP^j(\mathrm{\Sigma }C_g),SP^{j1}(\mathrm{\Sigma }C_g);๐
).$$
To see this, one simply observes that $`SP_n^k(C)`$ is open in $`SP^k(C)`$ which is a $`k`$ dimensional complex manifold and then one applies 4.3. We point out that such a spectral sequence was considered in the case of $`n=1`$ in (\[BCM\], theorem 4.1) and for $`g=0`$ and all $`n`$ in \[K3\]. The differentials as explained in the proof of 4.3 are induced from a cellular approximation of the maps
$$\nu :SP^j(C)SP^{(n+1)j}(C),x(n+1)x$$
$`4.7`$
More explicitly in this case, a chain complex for a Riemann surface $`C`$ of genus $`g`$ is given by $`2g`$ one dimensional classes (which we label $`e_1,\mathrm{},e_{2g}`$) and a two dimensional orientation class $`a=[C]`$. Now the homology of $`SP^{\mathrm{}}(C)`$ is generated as a ring by the symmetric products of these classes; i.e.
$$H_{}(SP^{\mathrm{}}(C))=E(e_1,\mathrm{},e_{2g})\mathrm{\Gamma }[a]$$
and $`H_{}(SP^n(C))`$ consists of all $`n`$-term products in the complex above (here $`E`$ is an exterior algebra and $`\mathrm{\Gamma }`$ is divided power algebra; see \[K1\] for details). It turns out that this homology embeds in $`C_{}(C)`$ and so one can think of the product of these classes as cells as well. We can investigate the boundary term 4.5 on these classes. The map $`\nu `$ is given by the composite $`C\genfrac{}{}{0pt}{}{\mathrm{\Delta }}{}C^{\times n}SP^n(C)`$. The primitive classes $`e_i`$ map to $`1\mathrm{}e_i\mathrm{}1`$ and hence map into $`H_{}(C)H_{}(SP^{n+1}(C))`$. For $`n>1`$ they clearly vanish in $`H_{}(SP^{n+1}(C),SP^n(C))`$ and so are not seen in the spectral sequence. The class $`a`$ on the other hand maps via $`\mathrm{\Delta }`$ to the class $`1\mathrm{}1C1\mathrm{}1+1\mathrm{}e_i\mathrm{}e_j\mathrm{}1`$ in $`H_2(C^{n+1})`$. The projection into $`H_{}(SP^{n+1}(C),SP^n(C))`$ vanishes if $`n>1`$ and is non trivial if $`n=1`$. More explicitly we have
Lemma 4.8: When $`n>1`$, all $`d^r`$ differentials vanish and the spectral sequence above collapses at the $`E^1`$ term. When $`n=1`$, there are higher differentials generated by $`d^1(1|a|)=2e_ie_{i+g}`$, $`1ig`$.
Example 4.9 ($`H_1(SP^k(C))`$): There are $`2g`$ one dimensional (torsion free) classes $`\stackrel{~}{e}_i`$ in $`H_1(SP_n^k(C))`$ corresponding to the classes $`e_ia^{k1}E_{k,0}^1`$ (which have dimension $`2(k1)+1`$). Also and when $`n=1`$, the class $`a^{k2}|a|`$ in $`E^1`$ gives rise to a generator in $`H_1(SP_1^k(C))`$ for all $`k>1`$. But $`|a|=2e_ie_{i+g}`$ (according to 4.8) and hence this is a 2-torsion class; i.e.
$$H_1(SP_n^k(C);๐)\{\genfrac{}{}{0pt}{}{๐^{2g},\text{when}n>1,k1}{๐_2๐^{2g},\text{when}k>n=1}$$
This also corresponds to $`H_1(\text{Map}_0^{}(C,๐^n))`$ as will be clear shortly (and as is expected according to 1.4!).
Example 4.10: There are inclusions $`H_{}(SP_n^k(C^{}))H_{}(SP_n^{k+1}(C^{}))`$ induced by a map of $`E^1`$ terms (in 4.6 above)
$$xyxay$$
As a corollary one can easily deduce the following standard fact (for another argument, see \[K3\] this volume): There are homology splittings (here $`SP^0(C^{})=`$)
$$H_{}(SP_n^k(C^{}))\underset{1ik}{}H_{}(SP_n^i(C^{}),SP_n^{i1}(C^{}))$$
ยง4.2 The homology of $`\text{Map}_0^{}(C,๐^n)`$
We now determine $`H_{}(\text{Map}^{}(C,๐^n);๐
)`$ for $`๐
=๐_2,๐_p`$ as a straightforward application of the calculations in ยง4.1. Recall that when $`g>0`$, the surface $`C_g`$ is described topologically by a 2-disc attached to a bouquet of $`2g`$ circles via the mapping which wraps around as a product of commutators. More precisely one has a cofibration sequence
$$S^1\genfrac{}{}{0pt}{}{f}{}\stackrel{2g}{}S^1\genfrac{}{}{0pt}{}{i}{}C_g\genfrac{}{}{0pt}{}{\pi }{}S^2$$
$`4.11`$
where $`i:^{2g}S^1C_g`$ is the one skeleton inclusion, and the map $`f`$ is given as a product of commutators $`[x_1,x_2][x_3,x_4]\mathrm{}[x_{2g1},x_{2g}]`$ (with the $`x_i`$โs denoting the generators of $`\pi _1(_{2g}S^1)=๐^{2g}`$). Applying the $`\text{Map}^{}(,๐^n)`$ functor to 4.11 yields the fibration
$$\mathrm{\Omega }^2(๐^n)\text{Map}^{}(C_g,๐^n)\mathrm{\Omega }(๐^n)^{2g}$$
$`4.12`$
We have $`H^{}(\mathrm{\Omega }(๐^n)^{2g})=H^{}(S^1)^{2g}H^{}(\mathrm{\Omega }^2S^{2n+1})`$. Note that 4.12 has simple coefficients (since the fiber is a loop space). Write again (cf. ยง2) $`H^{}(\mathrm{\Omega }^2S^3,๐_p)`$ as an exterior $`E(x_1,\mathrm{},x_{2p^{i+1}1},\mathrm{})`$ tensor a truncated algebra $`P_T(y_{2p2},\mathrm{},y_{2p^i2},\mathrm{})`$ where the $`x`$โs and $`y`$โs are generators in the stated dimensions.
Theorem 4.13: Assume $`g>1`$. Then the Serre spectral sequence for 4.12 collapses at $`E^2`$ when $`n>1`$. When $`n=1`$, the spectral sequence collapses with $`๐
_2`$ coefficients but has the mod-$`p`$ differentials ($`p>2`$)
$`d_{p^i}(x_{2p^i1})`$ $`=`$ $`{\displaystyle \frac{1}{p^i}}\left({\displaystyle \underset{1}{\overset{g}{}}}f_{2i1}f_{2i}\right)^{p^i}`$
$`d_{p^i}(y_{2p^{i+1}2})`$ $`=`$ $`\left[{\displaystyle \frac{1}{p^i}}\left({\displaystyle \underset{1}{\overset{g}{}}}f_{2i1}f_{2i}\right)^{(p1)p^i}\right]x_{2p^i1}.`$
where the $`f_i`$ are the one dimensional generators of $`\mathrm{\Omega }(๐^1)^{2g}(S^1)^{2g}\times (\mathrm{\Omega }S^3)^{2g}`$.
Proof: This is a counting argument. The homology splitting in 4.10 holds for $`k=\mathrm{}`$ and combining this with 1.7 (and field coefficients) we see that
$$H_{}(\text{Map}_k^{}(C,๐^n))\underset{k1}{}H_{}(SP_n^k(C^{}),SP_n^{k1}(C^{}))$$
We simply need identify generators. From 4.6, we read off the correspondence between generators in $`E^{\mathrm{}}`$ and generators in the serre spectral sequence as follows
$`e_i`$ $``$ $`f_i\text{and}|a|x_1`$
$`|e_i|^{}s`$ $``$ $`\text{generators of}H^2(\mathrm{\Omega }S^3)^{2g}`$
($`|e_i|,|a|E(2)`$ and $`e_iE(1)`$, but they propagate to $`E(k)`$ after multiplying by suitable powers of $`a`$ as indicated in 4.10). Now $`a`$ generates a divided power algebra in $`H_{}(SP^{\mathrm{}}(C))`$ and hence $`|a|`$ generates an $`H_{}(K(๐,3))`$ in $`H_{}(SP^{\mathrm{}}(\mathrm{\Sigma }C))`$ (as is known, the groups $`H_{}(K(๐,3))`$ and $`H^{}(\mathrm{\Omega }^2S^3)`$ are formally โdualโ to each other). More precisely, we can construct a correspondence between the two as follows (with mod-$`p`$ coefficients): let $`\gamma _i`$ be the divided power generators in $`\mathrm{\Gamma }(a)`$, then
$$|\gamma _{p^i}|x_i,|\gamma _{p^i}^{p1}|\gamma _p|y_i$$
The rest of the proof is now straightforward as the non-zero differentials for $`p>2`$ are entirely generated by the one in 4.8 (compare \[K1\]).
ยง5 Segalโs Stability Theorem
In this section we derive Segalโs result based on the spectral sequence in 4.3 which in this particular context takes the form
(5.1) There is a spectral sequence converging to $`H^{}(\text{Hol}_k(C_g,๐^n))`$, $`k2g1`$ with $`E^1`$ term
$${}_{h}{}^{}E_{}^{1}=\underset{\genfrac{}{}{0pt}{}{i+j=k}{i1}}{}E_{i,j}^1=\underset{\genfrac{}{}{0pt}{}{i+j=k}{i1}}{}H_{^{}}(LE_i;๐)H_{^{\prime \prime }}(SP^j(\mathrm{\Sigma }C_g),SP^{j1}(\mathrm{\Sigma }C_g);๐).$$
and identifiable $`d^1`$ differential. Here $`=2(kg)(n+1)+2g^{}^{\prime \prime }`$.
The terms $`LE_i`$ are constructed out of the Jacobi variety $`J`$ and its stratifications. Recall that by construction $`E_i`$ is the pull-back of the diagonal
$$\begin{array}{ccc}E_i& \genfrac{}{}{0pt}{}{}{}& SP^i(C_g)\times \mathrm{}\times SP^i(C_g)\\ & & \underset{n+1}{\underset{}{\mu \times \mathrm{}\times \mu }}\\ J& \genfrac{}{}{0pt}{}{\mathrm{\Delta }}{}& J(C_g)\times \mathrm{}\times J(C_g)\end{array}$$
$`5.2`$
The image under $`\mu `$ of $`SP^i(C)`$ in $`J(C)`$ is denoted by $`W_i`$. $`W_i`$ has complex dimensions $`i`$ and for for $`ig`$, $`\mu `$ is surjective and $`W_i=J`$.
We nee recall at this stage that the Abel-Jacobi map $`\mu :SP^i(C_g)J(C_g)`$ is an analytic fibration in the โstableโ range $`i2g1`$ with fiber $`๐^{ig}`$ (that the dimension of the fiber stabilizes is a consequence of Riemann-Roch. That $`\mu `$ is actually a fibration is a theorem of Mattuck \[Ma\]). So from 5.2 it is easy to deduce that for $`i2g`$
$$H_{}(LE_i,๐)H_{}(J(C_g),๐)H_{}(S^{2(ig)(n+1)};๐)$$
$`5.3`$
(i.e. $`LE_i`$ is obtained by first pulling back $`SP^i(C)^{n+1}`$ over $`J(C)`$ via $`\mathrm{\Delta }`$ and then collapsing the fat wedge in the fiber $`(๐^{ni})^{n+1}`$.)
Example 5.4: (the torsion free one-dimensional classes). Note that when $`i2g`$, we deduce from 5.3 that there are $`2g`$ classes in $`H_{}(LE_k)`$ of dimension $`2(kg)(n+1)+2g1`$ which yield $`2g`$ torsion free generators in $`H_1(\text{Hol}_k^{})`$.
Stable Classes and Segalโs theorem: When $`i2g1`$ (the stable range), the terms in $`{}_{h}{}^{}E_{i,j}^{1}`$ that survive to $`{}_{h}{}^{}E_{}^{\mathrm{}}`$ yield โstableโ classes in $`H^{}(\text{Hol}^{})`$ (and hence in $`H_{}(\text{Hol}^{})`$). The following is shown in (\[KM\], lemma 8.3)
Proposition 5.5: The induced map $`H_{}(\text{Hol}_k^{}(C_g,๐^n))H_{}(\text{Map}_k^{}(C_g,๐^n))`$ is an isomorphism on the stable classes.
Proof (sketch of alternate argument): When $`i2g1`$, $`E_i`$ is an $`(ig)(n+1)+2g`$ complex and maps to $`SP^{(ig)(n+1)+2g}(C)`$ (via a factoring of the map $`E_iSP^i(C)^{n+1}\genfrac{}{}{0pt}{}{+}{}SP^{i(n+1)}(C)`$ (where $`+`$ is concatenation). It follows that the stable terms in $`{}_{h}{}^{}E_{}^{1}`$ (corresponding to $`i2g1`$) map to $`E^1(N)`$ (see 4.6; here $`N=(kg)(n+1)+2g)`$). The main point is that in the stable range the spectral sequence in 5.1 and the one in 4.6 behave identically (same differentials described as in 4.8 above and \[KM\], lemma 7.8). The stable terms surviving to $`{}_{h}{}^{}E_{}^{\mathrm{}}`$ therefore map isomorphically to $`E^{\mathrm{}}(N)`$ and so the dual โstable classesโ in $`H_{}(\text{Hol}_k^{})`$ have isomorphic image in $`H_{}(SP_n^k(C))`$; i.e. in $`H_{}(SP_n^{\mathrm{}}(C))H_{}(\text{Map}_0^{}(C,๐^n)`$. Details omitted.
The unstable terms which appear when $`i<2g1`$ in the spectral sequence above form the interesting part and are harder to track down for their existence depends strongly on the geometry of the curve. It follows however that their contribution only starts appearing above a certain range and hence up to that range the homology of $`\text{Hol}_k(C_g,๐^n)`$ only consists of stable classes and by 5.5 must coincide with that of $`\text{Map}_k(C_g,๐^n))`$. This is exactly the essence of the stability theorem of Segal and we fill in the details below.
Proposition 5.6: $`H_{}(LE_i)=0`$ for $`>i(n+1)+2g`$ and for all $`1i2g`$.
Proof: Define
$$W_i^r=\{x\mu (SP^i(C_g))|\mu ^1(x)=๐^m,mr\}$$
A well-known theorem of Clifford asserts that in the range $`1i2g`$ the maximum of $`r`$ is $`i/2`$; that is $`W_i^r=\mathrm{}`$ for $`r>\frac{i}{2}`$. One can then filter $`\mu (SP^i(C))=W_i`$ by the descending filtration
$$W_i^{\frac{i}{2}}\mathrm{}W_i^1W_i$$
This leads to a spectral sequence converging to $`H_{}(LE_i)`$ with $`E^1`$ term
$$\underset{r}{\overset{\frac{i}{2}}{}}H_{}(W_i^r,W_i^{r+1})H_{}(S^{2r(n+1)})$$
Since the $`W_i^rW_i^{r+1}`$ are algebraic subvarieties of $`J(C_g)`$ (\[ACGH\]) we must have that $`H_{}(W_i^r,W_i^{r+1})=0`$ for $`>2g`$. This means that each term in the $`E^1`$ term above has no homology beyond $`2\frac{i}{2}(n+1)+2g`$ and hence
$$H_{}(LE_i)=0\text{for}>2\frac{i}{2}(n+1)+2g$$
as asserted. Note that when $`i=2g`$, we have that $`m`$ is uniformly $`ig=g=\frac{i}{2}`$ and the proposition is still valid in this case.
Corollary 5.7: (Segal) The inclusion
$$\text{Hol}_k(C_g,๐^n)\text{Map}_k(C_g,๐^n)$$
is a homology isomorphism up to dimension $`(k2g)(2n1)`$
Proof: Look in 4.3 at the unstable terms given by
$$H_{}(LE_i)H_{^{}}(SP^j(\mathrm{\Sigma }C_g),SP^{j1}(\mathrm{\Sigma }C_g)),i2g1$$
According to 5.6 these terms vanish for $`>i(n+1)+2g`$. By duality, they contribute therefore no homology to $`H_{}(\text{Hol}_k(C_g,๐^n);๐)`$ for
$$[2(kg)n+2k][i(n+1)+2g+3(ki)]$$
The expression on the left attains its minimum when $`i=2g`$ that is when
$$\left[2(kg)n+2k\right]\left[2g(n+1)+2g+3(k2g)\right]=(k2g)(2n1).$$
So when $`<(k2g)(2n+1)`$ the unstable terms do not contribute to $`H_{}(\text{Hol}_{k=i+j}(C_g,๐^n);๐)`$ and the proof now follows from proposition 5.5.
ยง6 The Fundamental Group
Since the attaching map of the two cell of $`C_g`$ is given by a product of commutators corresponding to Whitehead products (see 4.11), its suspension must be null-homotopic. This implies that $`\mathrm{\Sigma }C_g`$ splits as a wedge $`\mathrm{\Sigma }C_gS^2S^3`$ and one has
$$\mathrm{\Sigma }^iC_g\underset{1}{\overset{2g}{}}S^{i+1}S^{i+2},i1.$$
$`6.1`$
Lemma 6.2: All components of $`\text{Map}^{}(C_g,X)`$ are homotopy equivalent, and $`i>1`$, there is an isomorphism of homotopy groups
$$\pi _i(\text{Map}_0^{}(C_g,X))=\stackrel{2g}{}\pi _{i+1}(X)^{2g}\pi _{i+2}(X).$$
Proof: The statement about components is standard and we denote by $`\text{Map}_0^{}(C_g,X)`$ the component of null-homotopic maps. On the other hand and by definition $`\pi _i(\text{Map}_0^{}(C_g,X))=[S^iC_g,X]_{}`$. The splitting 6.1 yields (as sets) the bijection
$$[S^iC_g,X]_{}=[\underset{2g}{}S^{i+1}S^{i+2},X]_{}=\pi _{i+2}(X)\stackrel{2g}{}\pi _{i+1}(X)$$
In order for the decomposition above to induce a group homomorphism, it is necessary that 6.1 be a decomposition as โco-H spacesโ, that is that there is a map $`f_i:\mathrm{\Sigma }^iC_g\genfrac{}{}{0pt}{}{f_i}{}_1^{2g}S^{i+1}S^{i+2}`$ which commutes with the pinch maps up to homotopy. When $`i2`$, $`f_i=\mathrm{\Sigma }f_{i1}`$ is a suspension and hence is automatically a co-H map. The claim follows.
When $`i=1`$, the decomposition in 6.2 is not necessarily a group decomposition. The long exact sequence in homotopy associated to $`\mathrm{\Omega }^2X\text{Map}^{}(C_g,X)(\mathrm{\Omega }X)^{2g}`$ together with 6.2 indicate however that there is a short exact sequence of abelian groups
$$0\pi _3(X)\pi _1(\text{Map}_0^{}(C_g,X))\pi _2(X)^{2g}0$$
When $`X=๐^n`$, $`\pi _2(๐^n)๐`$ and $`\pi _3(๐^n)\pi _3(S^{2n+1})`$ which is zero when $`n>1`$. This shows that
$$\pi _1\text{Map}_0^{}(C,๐^n)=๐^{2g},\text{when}n>1$$
$`6.3(a)`$
When $`n=1`$, there is a short exact sequence
$$0๐\pi _1(\text{Map}_0^{}(C,๐^1))๐^{2g}0$$
$`6.3(b)`$
This is a central extension which turns out to be non-trivial as we now show.
First observe that any Riemann surface has meromorphic functions for sufficiently high degrees $`k`$ (this is certainly true when $`k2g`$ and we will assume this is the case throughout the section). Let $`f`$ be any such function and construct the map $`\text{Rat}_1(๐^1)\text{Hol}_k^{}(C,๐^1)`$ by post-composition with $`f`$. $`\text{Rat}_1(๐^1)`$ is the set of pairs of distinct points (a root and a pole) in $`๐`$ and has the homotopy type of $`S^1`$.
Lemma 6.4: The inclusion $`f^!:\text{Rat}_1(๐^1)\text{Hol}_k^{}(C,๐^1)`$ induces an injection at the level of $`\pi _1`$.
Proof: Observe that the diagram below commutes strictly
$$\begin{array}{ccc}\text{Rat}_1(๐^1)& \genfrac{}{}{0pt}{}{f^!}{}& \text{Hol}_k^{}(C,๐^1)\\ i_1& & i_2\\ \mathrm{\Omega }_1^2S^2& \genfrac{}{}{0pt}{}{g}{}& \text{Map}_k^{}(C,S^2)\end{array}$$
$`6.5`$
The vertical maps are inclusions. The bottom map on $`\pi _1`$ is an injection according to 6.3(b) and hence $`gi_1`$ is also and injection on $`\pi _1`$ (since $`i_1`$ on $`\pi _1`$ is an isomorphism between two copies of $`๐`$). The composition $`i_2f^!`$ is therefore injective on $`\pi _1`$ and hence so is $`f^!`$ as desired.
Next and from the description of $`\text{Hol}_k^{}`$ and $`\text{Div}_{k,k}(C)=SP^k(C)^2\mathrm{\Delta }`$ where $`\mathrm{\Delta }`$ is the generalized diagonal consisting of pairs of divisors with (at least) one point in common (see ยง3), we have the pullback diagram
$$\begin{array}{cccc}\text{Hol}_k^{}(C,๐^1)& \genfrac{}{}{0pt}{}{\varphi }{}& \text{Div}_{k,k}(C)& \\ \pi & & \mu ^2& \\ J(C)& \genfrac{}{}{0pt}{}{\mathrm{\Delta }}{}& J(C)^2& \end{array}$$
$`6.6`$
where $`\pi :\text{Hol}_k^{}(C,๐^n)\genfrac{}{}{0pt}{}{p}{}SP^k(C)\genfrac{}{}{0pt}{}{\mu }{}J(C)`$ is the map that sends a holomorphic map to the equivalence class of its divisor of zeros and $`\mathrm{\Delta }`$ is the diagonal. The top horizontal map $`\varphi `$ is an inclusion. We denote by $`F`$ the homotopy fiber of $`\mu ^2`$ and $`\pi `$.
Lemma 6.7: The map $`p:\text{Hol}_k^{}(C,๐^1)SP^k(C)`$ is surjective at the level of fundamental groups.
Proof: For $`k2g1`$ the restriction of the Mattuck fibration (cf. ยง5) to $`SP^k(C)`$ becomes a complex vector bundle
$$๐^{kg}SP^k(C)\genfrac{}{}{0pt}{}{\mu }{}J(C)$$
(i.e. the set of linearly equivalent divisors avoiding a point is a hyperplane in $`๐^{kg}`$). It follows that $`\mu _{}`$ is an isomorphism on $`\pi _1(SP^k(C))`$. Since the fiber of $`\pi `$ is connected (see 6.10), it follows that $`\pi `$ is surjective at the level of $`\pi _1`$ and hence is $`p`$.
Theorem 6.8: Suppose $`k2g`$. Then $`\pi _1(\text{Hol}_k^{}(C,๐^1))`$ is generated (multiplicatively) by classes $`e_1,\mathrm{},e_{2g}`$ and $`\tau `$ such that the commutators
$$[e_i,e_{g+i}]=\tau ^2$$
and all other commutators are zero.
We will prove this in a series of lemmas. To begin with, lemmas 6.4 and 6.7 show that there is a sequence
$$0๐\genfrac{}{}{0pt}{}{f_{}}{}\pi _1\text{Hol}_k^{}(C,๐^1)\genfrac{}{}{0pt}{}{p_{}}{}๐^{2g}0$$
$`6.9`$
which is exact at both ends. In 6.13 we show that this sequence maps to 6.3(b) and hence Im($`f_{})`$ Ker $`(p_{})`$. That Ker$`(p)`$Im$`(f_{})`$ follows from 6.11 and the discussion next.
Lemma 6.10: $`\pi _1(\text{Div}_{k,k}(C)),k>1`$ has torsion free generators $`a_i,b_j,1i,j2g`$ and $`\tau `$. The $`a_i`$ (resp. $`b_i`$) are represented by a root (resp. a pole) going around the generators in a symplectic basis of $`H_1(C)`$, while $`\tau `$ is given by a zero (or pole) going around a pole (resp. a zero).
Proof: There are a few ways to see that there is a short exact sequence
$$0๐\pi _1(\text{Div}_{k,k}(C))๐^{4g}0$$
with generators having the desired properties. $``$ By considering the map $`\mu ^2`$ in 6.6 and analyzing the fiber $`F`$ given by the complement of a hyperplane in $`๐^{kg}\times ๐^{kg}=๐^{2(kg)}`$ (the fiber of $`\mu `$ restricted to $`SP^k(C)`$ is $`๐^{kg}`$ as indicated in 6.7). Here $`\pi _1(F)=๐`$. $``$ Mimicking the proof of Jones for the case $`C=๐^1`$. One considers the subspace $`U`$ of $`\text{Div}_{k,k}(C))`$ consisting of pairs $`(D_0,D_1)`$ with the roots of $`D_1`$ all distinct. This is the complement of a (real) codimension $`2`$ subspace and there is a surjection $`\pi _1(U)\pi _1(\text{Div}_{k,k}(C))`$. Now there is a fibration $`SP^k(CQ_{k+1})\pi _1(U)\pi _1(F(C,k)`$ with section (where $`Q_{k+1}`$ is a set of $`k+1`$ distinct points). Since $`\pi _1(SP^k(CQ_{k+1}))=H_1(CQ_{k+1})=๐^{2g}\times ๐^k`$, there is a semi-direct product of $`๐^{2g}\times ๐^k`$ with the ordered braid group $`B_k(C)`$ with quotient at least $`๐^{4g}`$. A similar argument as in \[S\], lemma 6.4 implies the answer.
Lemma 6.11: $`\pi _1(\text{Hol}_k^{}(C,๐^1))`$ has generators $`\tau `$ and $`e_i,1ig`$ which map under $`\varphi _{}:\pi _1(\text{Hol}_k^{}(C,๐^1))\pi _1(\text{Div}_{k,k}(C))`$ as follows: $`\varphi _{}(\tau )=\tau `$ (the generator of the same name) and $`\varphi _{}(e_i)=a_ib_i`$. In particular $`\varphi _{}`$ is an injection.
Proof: Diagram 6.6 induces a map of fibrations and since $`\pi _2(J)=0`$, a map of short exact sequences
$$\begin{array}{ccccccccc}0& & ๐& & \pi _1(\text{Hol}_k^{}(C,๐^1))& & ๐^{2g}& 0& \\ & & =& & \varphi _{}& & \mathrm{\Delta }_{}& & \\ 0& & ๐& & \pi _1(\text{Div}_{k,k}(C))& & ๐^{4g}& 0& \end{array}$$
The map $`\mathrm{\Delta }_{}`$ is an injection and then so is the middle map. The generator $`\tau \pi _1(\text{Hol}_k^{})`$ constructed in 6.4 corresponds exactly to the generator coming from $`\pi _1(F)`$ and is described by zeros linking with poles. The rest of the claim follows from the commutativity of the above diagram.
Proof of theorem 6.8: The generators of $`\pi _1`$ and their images under $`\varphi `$ having been determined, we next look at the image of the commutators
$$\varphi ([e_i,e_j])=a_i\underset{}{b_ia_j}b_jb_i^1\underset{}{a_i^1b_j^1}a_j^1$$
Suppose we know that for $`k2`$ and $`1i,jg`$
$$a_ib_ja_i^1b_j^1=\{\genfrac{}{}{0pt}{}{\tau ,|ij|=g}{1,\text{otherwise}}\pi _1(\text{Div}_{k,k}(C)),$$
$`6.12`$
then $`\varphi [e_i,e_j]`$ becomes (after transforming the underbraced terms) and for $`|ij|=g`$
$`\varphi ([e_i,e_j])`$ $`=`$ $`a_i\tau a_jb_ib_jb_i^1b_j^1a_i^1\tau a_j^1`$
$`=`$ $`\tau ^2a_ja_i[b_i,b_j]a_i^1a_j^1`$
$`=`$ $`\tau ^2[a_i,a_j]=\tau ^2`$
This is the claim in 6.8 and therefore the theorem would follow as soon as we prove 6.12.
Consider the sub-divisor space $`\text{Div}_{k,1}^2(C))\text{Div}_{k,k}(C))`$ consisting of pairs $`(D_1,q)SP^k(C)\times (C)`$ with the property that $`qD_1`$ (here again $`k>1`$). There is a projection and a fibration
$$SP^k(C\{,q\})\text{Div}_{k,1}(C)(C)$$
The generators $`b_i\pi _1(C)`$ act on the fundamental group of the fiber $`\pi _1(SP^k(C\{,q\})`$ and this action describes the commutator map in 6.12. Since $`\pi _1(SP^k(C\{,q\})`$ is abelian, we can look at the action in homology. This has been already determined in (\[C<sup>2</sup>M<sup>2</sup>\], prop. 10.12) and is exactly given by 6.12 ($`b_i`$ is $`\tau _{1i}`$ in their notation). This completes the proof.
Proposition 6.13: For $`k2g`$ and all $`n1`$, we have an isomorphism
$$\pi _1(\text{Hol}_k^{}(C,๐^n))\pi _1(\text{Map}_0^{}(C,๐^n))$$
Proof: When $`n>1`$, one can show that $`\pi _1(\text{Hol}_k^{}(C,๐^n))`$ is abelian (this has been proven in \[Ki\], lemma 3.5). The equivalence in $`\pi _1`$ when $`n>1`$ becomes an equivalence at the level of $`H_1`$ which we know is true by Segalโs theorem. In fact when $`n>1`$, $`\pi _1(\text{Hol}_k^{}(C,๐^n))=\pi _1(\text{Map}_k^{}(C,๐^n))๐^{2g}`$ by 6.3(a).
Suppose $`n=1`$. The proof of 6.11 shows that the sequence in 6.9 is actually exact. We would like to see that 6.9 maps to the exact sequence in 6.3(b). The essential point is to show that the following homotopy commutes
$$\begin{array}{cccc}\text{Hol}_k^{}(C,๐^1)& \genfrac{}{}{0pt}{}{p}{}& SP^{\mathrm{}}(C)& \\ i& & & \\ \text{Map}_k^{}(C,๐^1)& (\mathrm{\Omega }S^2)^{2g}& (\mathrm{\Omega }๐^{\mathrm{}})^{2g}& \end{array}$$
where again $`p`$ sends a holomorphic map to the divisor of its zeros, the bottom composite $`g:\text{Map}_k^{}(C,๐^1)(\mathrm{\Omega }๐^{\mathrm{}})^{2g}`$ is restriction to the one skeleton followed by postcomposition with $`S^2๐^{\mathrm{}}`$. Since both spaces on the far right are Eilenberg-Maclane spaces $`(S^1)^g`$, one simply need consider the effect of $`p`$ and $`gi`$ on cohomology (recall that $`i`$ is an isomorphism on $`H^1`$ by 1.3). Label generators of $`H^1((S^1)^{2g})`$ by $`f_i,1i2g`$. As is explicit in theorem 4.13, the $`g^{}(f_i)`$ correspond to the non-trivial torsion free classes in $`H^1(\text{Map}_k^{})`$ (see 4.13 and 4.9). Similarly the construction of the stable classes in ยง5 (see example 5.4) implies precisely that $`p^{}(f_i)`$โs are the torsion free generators of $`H^1(\text{Hol}_k^{})`$. That is $`p^{}=i^{}g^{}`$ and hence $`pig`$.
Now and upon applying $`\pi _1`$ to the terms of the diagram above as well as to those in 6.5 we get a map from 6.9 down to 6.3(b). Since the end terms of the sequences are isomorphic, it follows by the five lemma that the middle terms are isomorphic; i.e. $`\pi _1(\text{Hol}_k^{})\pi _1(\text{Map}_k^{})`$.
Corollary 6.14: $`\pi _1(\text{Map}_0^{}(C,๐^n))`$ is generated by classes $`e_1,\mathrm{},e_{2g}`$ and $`\alpha `$ such that the commutators $`[e_i,e_{g+i}]=\alpha ^2`$ and all other commutators are zero.
Note: Proposition 6.13 is not sufficient to imply that Segalโs homology equivalence 5.7 can be upgraded to a homotopy equivalence. A detailed study of the actions of $`\pi _1`$ on the universal covers is needed.
ยง7 Spaces of Unbased Maps
There is an important distinction between $`\text{Map}^{}(C,X)`$ and $`\text{Map}(C,X)`$. For one thing, while the topology of the based mapping space doesnโt vary with the degree, this is no longer true in the unbased case. Let $`\text{Map}_f(C_g,X)`$ denote the component containing a given map $`f`$, $`x_0C_g`$ the basepoint and consider again the โevaluationโ fibration
$$\text{Map}_f^{}(C_g,X)\text{Map}_f(C_g,X)\genfrac{}{}{0pt}{}{ev}{}X,ev(g)=g(x_0)$$
$`7.1`$
Associated to 7.1 is a long exact sequence of homotopy groups
$$\pi _i\text{Map}_f^{}(C_g,X)\pi _i\text{Map}_f(C_g,X)\pi _i(X)\genfrac{}{}{0pt}{}{}{}\pi _{i1}\text{Map}_f^{}(C_g,X)$$
Suppose $`X`$ is simply connected. Notice that given a map $`f:C_gX`$, $`f`$ is null on $`S^1`$ and hence factors (up to homotopy) as in
$$f:C_gS^2\genfrac{}{}{0pt}{}{\overline{f}}{}X$$
One then has the following diagram of vertical fibrations over $`X`$
$$\begin{array}{ccccc}\mathrm{\Omega }_{\overline{f}}^2X& & \text{Map}_f^{}(C_g,X)& & (\mathrm{\Omega }X)^{2g}\\ & & & & \\ _{\overline{f}}^2X& & \text{Map}_f(C_g,X)& & \text{Map}(S^1,X)\\ ev& & ev& & ev\\ X& \genfrac{}{}{0pt}{}{=}{}& X& \genfrac{}{}{0pt}{}{=}{}& X\end{array}$$
$`7.3`$
inducing a diagram of homotopy groups and boundary homomorphisms
$$\begin{array}{ccccc}\pi _i(X)& \genfrac{}{}{0pt}{}{=}{}& \pi _i(X)& \genfrac{}{}{0pt}{}{=}{}& \pi _i(X)\\ _1& & & & _2\\ \genfrac{}{}{0pt}{}{\pi _{i1}(\mathrm{\Omega }^2(X))}{=\pi _{i+1}(X)}& & \pi _{i1}(\text{Map}_f^{}(C_g,X))& & \genfrac{}{}{0pt}{}{\pi _{i1}(\mathrm{\Omega }_{fi}X)^{2g})}{=\pi _i(X)^{2g}}\end{array}$$
Proposition 7.4: Assume $`X`$ simply connected, $`f:C_gX`$ and $`\overline{f}`$ such that $`C_g\genfrac{}{}{0pt}{}{\pi }{}S^2\genfrac{}{}{0pt}{}{\overline{f}}{}X`$ (up to homotopy). Assume $`i>2`$. Then the homotopy boundary
$$:\pi _i(X)\pi _{i1}(\text{Map}_f^{}(C_g,X))=\pi _i(X)^{2g}\pi _{i+1}(X)$$
for the fibration $`\text{Map}_f^{}(C_g,X)\text{Map}_f(C_g,X)\genfrac{}{}{0pt}{}{ev}{}X`$ factors through the term $`\pi _{i+1}(X)`$ and is given (up to sign) by the Whitehead product: $`\alpha =[\alpha ,\overline{f}].`$
Proof: The sequence $`0\pi _{i+1}(X)\pi _{i1}(\text{Map}_f^{}(C_g,X))\pi _i(X)^{2g}0`$ splits canonically (according to 6.2) and with respect to that splitting one has $`=_1+_2`$. By Whiteheadโs theorem we have that $`_1\alpha =[\alpha ,\overline{f}]`$ while $`_2\alpha =0`$. The proof follows.
Remark: When $`i=2`$ the situation has to be handled differently for $`\pi _1(\text{Map}_f^{}(C_g,X))`$ does not necessarily split as in 7.4.
Lemma 7.5: $`\pi _1(\text{Map}_d(C,๐^n))๐^{2g}`$ when $`n>1`$.
Proof: Consider the Hopf fibration $`S^{2n+1}๐^n๐^{\mathrm{}}`$ and the induced fibration
$$\text{Map}(C,S^{2n+1})\text{Map}_d(C,๐^n)\text{Map}_d(C,๐^{\mathrm{}})(S^1)^{2g}\times ๐^{\mathrm{}}$$
Since $`\pi _1(\text{Map}(C,S^{2n+1}))=0`$ when $`n>1`$, the result follows.
We next address the case $`n=1`$ and analyze 7.3 when $`X=๐^1`$. Let $`f:S^2X`$ be a degree $`d`$ map. The boundary term in the left vertical fibration $`\pi _1(\mathrm{\Omega }_f^2X)\genfrac{}{}{0pt}{}{}{}\pi _1(_f^2(X))`$ is given according to 7.4 by multiplication by $`2d`$. On the other hand, the right vertical fibration has a section and hence we get the diagram of fundamental groups
$$\begin{array}{ccccccccc}0& & ๐& & \pi _1(\text{Map}_d^{})& & ๐^{2g}& & 0\\ & & & & & & =& & \\ 0& & ๐_{2d}& & \pi _1(\text{Map}_d)& & ๐^{2g}& & 0\end{array}$$
Corollary 7.6: (Larmore-Thomas) $`\pi _1(\text{Map}_d(C,S^2))`$ is generated by classes $`e_1,\mathrm{},e_{2g}`$ and $`\alpha `$ such that
$$\alpha ^{2|d|}=1,[e_i,e_{g+i}]=\alpha ^2$$
All other commutators are zero. In particular $`\text{Map}_d(C,๐^1)`$ and $`\text{Map}_d^{}(C,๐^1)`$ have different homotopy types whenever $`d\pm d^{}`$.
Corollary 7.7: For $`d2g`$, $`\pi _1(\text{Hol}_d(C,๐^1))`$ is generated by $`e_1,\mathrm{},e_{2g}`$ and $`\alpha `$ with $`\alpha ^{2|d|}=1,[e_i,e_{g+i}]=\alpha ^2`$ and all other commutators are zero.
Proof: This is equivalent to showing that $`\pi _1(\text{Hol}_d(C,๐^1))`$ corresponds to $`\pi _1(\text{Map}_d(C,S^2))`$ in that range but this follows from a direct comparison of the evaluation fibrations for both mapping spaces and theorem 1.4.
In the case of maps into $`๐^n`$, $`n2`$, the fundamental group is not enough to distinguish between connected components. Proposition 7.4 still implies
Proposition 7.8: $`\text{Map}_k(C_g,๐^{2d})`$ and $`\text{Map}_l(C_g,๐^{2d})`$ have different homotopy types whenever $`lk(2)`$.
Proof: The long exact sequence for the evaluation fibration gives again
$$\mathrm{}\pi _{2d+1}๐^{2d}\genfrac{}{}{0pt}{}{}{}\pi _{2d}(\text{Map}_k^{})\pi _{2d}(\text{Map}_k)\pi _{2d}(๐^{2d})\genfrac{}{}{0pt}{}{}{}\mathrm{}$$
where $`\text{Map}_k^{}`$ stands for $`\text{Map}^{}(C_g,๐^{2d})`$. According to 7.4 the sequence above becomes
$$\pi _{2d+1}๐^{2d}=๐\eta \genfrac{}{}{0pt}{}{k[\iota ,\eta ]}{}\pi _{2d+2}(๐^{2d})=๐_2\pi _{2d}(\text{Map}_k)0$$
where $`\iota `$ is the generator of $`\pi _2(๐^{2d})`$. The Whitehead product pairing for complex projective spaces
$$\pi _{2m+1}(๐^m)\pi _2(๐^m)\pi _{2m+2}(๐^m)$$
is worked out in \[P\]. The result there is that the Whitehead product is zero if $`m`$ is odd and non-zero if $`m`$ is even. Since we are in the case $`m=2d`$, the map $`๐\eta \genfrac{}{}{0pt}{}{k[\iota ,\eta ]}{}๐_2`$ is in fact multiplication by $`k`$ and the proposition follows right away. |
warning/0003/hep-th0003167.html | ar5iv | text | # Renormalisability of the SU(2)รU(1) Electroweak Theory with Massive W Z Fields and Massive Matter Fields
## Abstract
We extend the previous work and study the renormalisability of the SU<sub>L</sub>(2) $`\times `$ U<sub>Y</sub>(1) electroweak theory with massive W Z fields and massive matter fields. We expound that with the constraint conditions caused by the W Z mass term and the additional condition chosen by us we can still performed the quantization in the same way as before. We also show that when the $`\delta `$ functions appearing in the path integral of the Green functions and representing the constraint conditions are rewritten as Fourier integrals with Lagrange multipliers $`\lambda _a`$ and $`\lambda _y`$, the total effective action consisting of the Lagrange multipliers, ghost fields and the original fields is BRST invariant. Furthermore, with the help of the the renormalisability of the theory without the the mass term of matter fields, we find the general form of the divergent part of the generating functional for the regular vertex functions and prove the renormalisability of the theory with the mass terms of the W Z fields and the matter fields.
PACS numbers: 03.65.Db, 03.80.+r, 11.20.Dj
I. Introduction
Owing to the lack of experimental evidence for the Higgs Bosons and to the unsatisfying treatment on quantization the non-Abelian theory with massive gauge fields has been reinvestigated \[1-9\]. Particularly, it has been clarified \[1-3\] that the SU(n) theory with massive gauge fields and the SU(2)$`\times `$U(1) theory of S.L.Glashow with massive W Z fields are renormalisable. In the present paper we will extend these work and study the renormalisability of the latter electroweak theory with the massive W Z fields and massive matter fields. For the sake of convenience we assume that the matter fields consist only of the electron and electron-neutrino fields.
With the W Z mass term and the matter field mass term directly added to the Lagrangian by hand, the classical equations of motion will yield complicated constraint conditions containing products of the field functions. Moreover, when the constraints are expressed so that the gauge field parts contain no mass parameters the matter field parts will have a negative dimension coefficient $`m/M^2`$, where $`M`$ and $`m`$ are the mass parameters of the W fields and the electron fields respectively.
As in the case of Ref. , since the mass term is invariant under an infinitesimal gauge transformation with $`\delta \theta _1`$ and $`\delta \theta _2`$ equal to zero and $`\delta \theta _3`$ equal to $`\delta \theta _y`$, where $`\theta _a`$ and $`\theta _1`$ are the parameters of the gauge group, an additional constraint condition should be properly chosen. We will expound that with the constraint conditions caused by the mass term and the additional condition chosen by us we can performed the quantization and construct the ghost action in a way similar to that used in Refs. . We will also show that when the $`\delta `$ functions appearing in the path integral of the Green functions and representing the constraint conditions are rewritten as Fourier integrals with Lagrange multipliers $`\lambda _a`$ and $`\lambda _y`$, the total effective action consisting of the Lagrange multipliers, ghost fields and the original fields is BRST invariant. A special thing is that the effective action has a matterโghost term coming from the matter field parts of the constraint conditions and containing the factor $`m/M^2`$.
We will follow the procedure of Ref. and use the generalized form of the theory containing $`\lambda _a`$, $`\lambda _y`$ and their sources in the generating functional for the Green functions to study the renormalisability of the theory containing only the original fields and the ghost fields. Namely, after deriving the SlavnovโTaylor identities and the additional identities for the generating functional $`\mathrm{\Gamma }`$ for the regular vertex functions with the help of the generalized form of the theory, we will let vanish the functional derivatives of $`\mathrm{\Gamma }`$ with respect to the classical fields of these Lagrange multipliers. In this way the divergent part of $`\mathrm{\Gamma }`$ will be shown to satisfy a set of equations which can still be treated. Furthermore, with the help of the the renormalisability of the theory without the the mass term of matter fields, we will be able to find the general form of the divergent part of $`\mathrm{\Gamma }`$ and prove that the mass term of the matter fields is also harmless to the renormalisability of the theory.
In spite of the extra complexeity caused by the mass term of the matter fields we will write this paper in the similar form as that of Ref. . In section $`2`$ we will find the constraint conditions coming from the W Z mass term and choose the additional constraint condition. The method of quantization will be explained in section $`3`$. Setion $`4`$ is devoted to prove the renormalisability of the theory. Concluding remarks will be given in the final section.
II. Original and Additional Constraint Conditions
The matter fields will be often denoted by $`\psi (x)`$ and $`\overline{\psi }(x)`$ and they only contain the electron fields and electron-neutrino fields in the present work. The former stands for the purely left-handed neutrino field $`\nu _L`$, the left- and right-handed parts of the electron field namely $`e_L`$, $`e_R`$, and the latter stands for $`\overline{\nu }_L`$, $`\overline{e}_L`$ and $`\overline{e}_R`$. Therefore the mass term of the matter fields is
$`_{\psi m}(x)=m\overline{e}_L(x)e_R(x)m\overline{e}_R(x)e_L(x).`$ (2.1)
Next let $`W_{a\mu }(x)`$, $`W_{y\mu }(x)`$ be the SU<sub>L</sub>(2) and U<sub>Y</sub>(1) gauge fields and $`g`$, $`g_1`$ be the coupling constants. Thus the W Z mass term in the Lagrangian is
$`_{WM}={\displaystyle \frac{1}{2}}M^2W_{a\mu }W_a^\mu +{\displaystyle \frac{1}{2}}M^2\left({\displaystyle \frac{g_1}{g}}\right)^2W_{y\mu }W_y^\mu M^2\left({\displaystyle \frac{g_1}{g}}\right)W_{3\mu }W_y^\mu ,`$ (2.2)
or
$$_{WM}=\frac{1}{2}M^2W_{1\mu }(x)W_1^\mu (x)+\frac{1}{2}M^2W_{2\mu }(x)W_2^\mu (x)+\frac{1}{2}M_z^2Z_\mu (x)Z^\mu (x),$$
where $`M_z^2`$ stands for $`g^2(g^2+g_1^2)M^2`$, and $`Z_\mu (x)`$, $`A_\mu (x)`$ are the field functions of Z boson and photon, namely
$`Z_\mu ={\displaystyle \frac{1}{\sqrt{(g^2+g_1^2)}}}(gW_{3\mu }g_1W_{y\mu }),`$ (2.3)
$`A_\mu ={\displaystyle \frac{1}{\sqrt{(g^2+g_1^2)}}}\epsilon (g_1W_{3\mu }+gW_{y\mu }),`$ (2.4)
where $`\epsilon `$ is $`1`$ or $`1`$.
The original Lagrangian of the SU<sub>L</sub>(2) $`\times `$ U<sub>Y</sub>(1) electroweak theory with the mass term $`_{WM}`$ is
$`=_{\psi m}+_\psi +_{\psi W}+_{WM}+_{WL}+_{WY},`$ (2.5)
where $`_\psi `$ describe the pure matter fields, $`_{\psi W}`$ is the coupling term between the matter and gauge fields. $`_{WL}`$ and $`_{WY}`$ are the gauge field parts without mass terms, namely
$`_{WL}={\displaystyle \frac{1}{4}}F_{a\mu \nu }F_a^{\mu \nu },`$ (2.6)
$`_{WY}={\displaystyle \frac{1}{4}}B_{\mu \nu }B^{\mu \nu },`$ (2.7)
where
$`F_{a\mu \nu }=_\mu W_{a\nu }_\nu W_{a\mu }gC_{abc}W_{b\mu }W_{c\nu },`$ (2.8)
$`B_{\mu \nu }=_\mu W_{y\nu }_\nu W_{y\mu }.`$ (2.9)
$`C_{abc}`$ stands for the structure constants of SU<sub>L</sub>(2) with $`C_{123}`$ equal to $`1`$.
Denote by $`\theta _a(x),\theta _y(x)`$ the parameters of the gauge group. Thus, under an infinitesimal gauge transformation, the fields $`W_a^\mu `$, $`W_y^\mu `$, $`\psi `$ and $`\overline{\psi }`$ transform as
$`\delta W_a^\mu (x)={\displaystyle \frac{1}{g}}^\mu \delta \theta _a(x)C_{abc}W_c^\mu (x)\delta \theta _b(x),`$
$`\delta W_y^\mu (x)={\displaystyle \frac{1}{g_1}}^\mu \delta \theta _y(x),`$
$`\delta \nu _L(x)={\displaystyle \frac{i}{2}}\delta \theta _1(x)e_L(x)+{\displaystyle \frac{1}{2}}\delta \theta _2(x)e_L(x)+{\displaystyle \frac{i}{2}}\delta \theta _3(x)\nu _L(x){\displaystyle \frac{i}{2}}\delta \theta _y(x)\nu _L(x),`$
$`\delta e_L(x)={\displaystyle \frac{i}{2}}\delta \theta _1(x)\nu _L(x){\displaystyle \frac{1}{2}}\delta \theta _2(x)\nu _L(x){\displaystyle \frac{i}{2}}\delta \theta _3(x)e_L(x){\displaystyle \frac{i}{2}}\delta \theta _y(x)e_L(x),`$
$`\delta e_R(x)=i\delta \theta _y(x)e_R(x),`$
$`\delta \overline{\nu }_L(x)={\displaystyle \frac{i}{2}}\delta \theta _1(x)\overline{e}_L(x)+{\displaystyle \frac{1}{2}}\delta \theta _2(x)\overline{e}_L(x){\displaystyle \frac{i}{2}}\delta \theta _3(x)\overline{\nu }_L(x)+{\displaystyle \frac{i}{2}}\delta \theta _y(x)\overline{\nu }_L(x),`$
$`\delta \overline{e}_L(x)={\displaystyle \frac{i}{2}}\delta \theta _1(x)\overline{\nu }_L(x){\displaystyle \frac{1}{2}}\delta \theta _2(x)\overline{\nu }_L(x)+{\displaystyle \frac{i}{2}}\delta \theta _3(x)\overline{e}_L(x)+{\displaystyle \frac{i}{2}}\delta \theta _y(x)\overline{e}_L(x),`$
$`\delta \overline{e}_R(x)=i\delta \theta _y(x)\overline{e}_R(x).`$
$`\delta _{\psi m}`$ can be written as
$`\delta _{\psi m}=f_a(x)\delta \theta _a(x)+f_y(x)\delta \theta _y(x),`$ (2.10)
where
$`f_1(x)={\displaystyle \frac{i}{2}}m\left\{\overline{\nu }_L(x)e_R(x)\overline{e}_R(x)\nu _L(x)\right\},`$ (2.11)
$`f_2(x)={\displaystyle \frac{1}{2}}m\left\{\overline{\nu }_L(x)e_R(x)+\overline{e}_R(x)\nu _L(x)\right\},`$ (2.12)
$`f_3(x)={\displaystyle \frac{i}{2}}m\left\{\overline{e}_R(x)e_L(x)\overline{e}_L(x)e_R(x)\right\},`$ (2.13)
$`f_y(x)=f_3(x).`$ (2.14)
Therefore the action transforms as
$`\delta {\displaystyle d^4x(x)}=\delta {\displaystyle d^4x\left\{_{WM}(x)+_{\psi m}(x)\right\}}`$
$`={\displaystyle }d^4x\{({\displaystyle \frac{M^2}{g}}_\mu W_1^\mu (x)+{\displaystyle \frac{M^2}{g}}g_1W_{2\mu }(x)W_y^\mu (x)+f_1(x))\delta \theta _1`$
$`+\left({\displaystyle \frac{M^2}{g}}_\mu W_2^\mu (x){\displaystyle \frac{M^2}{g}}g_1W_{1\mu }(x)W_y^\mu (x)+f_2(x)\right)\delta \theta _2`$
$`+({\displaystyle \frac{M^2}{g}}_\mu W_3^\mu (x){\displaystyle \frac{M^2}{g^2}}g_1_\mu W_y^\mu (x)+f_3(x))(\delta \theta _3\delta \theta _y)\}.`$ (2.15)
Since the classical equations of motion make the action invariant under an arbitrary infinitesimal transformation of the field functions, they certainly make the mass term invariant under an arbitrary infinitesimal gauge transformation. This means that when $`M`$ is not equal to zero, the classical equations of motion leads to the following constraint conditions
$`{\displaystyle \frac{M^2}{g}}_\mu W_1^\mu (x)+{\displaystyle \frac{M^2}{g}}g_1W_{2\mu }(x)W_y^\mu (x)+f_1(x)=0,`$ (2.16)
$`{\displaystyle \frac{M^2}{g}}_\mu W_2^\mu (x){\displaystyle \frac{M^2}{g}}g_1W_{1\mu }(x)W_y^\mu (x)+f_2(x)=0,`$ (2.17)
$`{\displaystyle \frac{M^2}{g}}_\mu W_3^\mu (x){\displaystyle \frac{M^2}{g^2}}g_1_\mu W_y^\mu (x)+f_3(x)=0.`$ (2.18)
These are the original constraint conditions. Since the mass term is invariant under an infinitesimal gauge transformation with $`\delta \theta _1`$ and $`\delta \theta _2`$ equal to zero and $`\delta \theta _3`$ equal to $`\delta \theta _y`$, $`_\mu W_3^\mu `$ and $`_\mu W_y^\mu `$ appear in one constraint. We now choose an additional condition and replace (2.18) with
$`{\displaystyle \frac{M^2}{g}}_\mu W_3^\mu (x)+{\displaystyle \frac{M^2}{g}}g_1W_{3\mu }(x)W_y^\mu (x)+f_3(x)=0,`$ (2.19)
$`_\mu W_y^\mu (x)+gW_{3\mu }(x)W_y^\mu (x)=0.`$ (2.20)
III. Quantization and BRST Invariance
Write (2.16), (2.17) and (2.19),(2.20) as
$`\mathrm{\Phi }_a(x)=0,\mathrm{\Phi }_y(x)=0,`$ (3.1)
with
$`\mathrm{\Phi }_1(x)=_\mu W_1^\mu (x)+g_1W_{2\mu }(x)W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}f_1(x),`$ (3.2)
$`\mathrm{\Phi }_2(x)=_\mu W_2^\mu (x)g_1W_{1\mu }(x)W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}f_2(x),`$ (3.3)
$`\mathrm{\Phi }_3(x)=_\mu W_3^\mu (x)+g_1W_{3\mu }(x)W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}f_3(x),`$ (3.4)
$`\mathrm{\Phi }_y(x)=_\mu W_y^\mu (x)+gW_{3\mu }(x)W_y^\mu (x).`$ (3.5)
Taking the constraint conditions (3.1) into account one should write the path integral of the Green functions inolving only the original fields as
$`{\displaystyle \frac{1}{N_0}}{\displaystyle ๐[๐ฒ,\overline{\psi },\psi ]\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]\underset{a^{},x^{}}{}\delta \left(\mathrm{\Phi }_a^{}(x^{})\right)\delta \left(\mathrm{\Phi }_y(x^{})\right)W_{a\mu }(x)W_{b\nu }(y)\mathrm{}\mathrm{exp}\{\mathrm{i}I\}},`$ (3.6)
where
$`I={\displaystyle d^4x(x)},`$
$`N_0={\displaystyle ๐[๐ฒ,\overline{\psi },\psi ]\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]\underset{a^{},x^{}}{}\delta \left(\mathrm{\Phi }_a^{}(x^{})\right)\delta \left(\mathrm{\Phi }_y(x^{})\right)\mathrm{exp}\{\mathrm{i}I\}}.`$
Since only the field functions which satisfy the constraint conditions can play roles in the integral (3.6), the value of the Lagrangian can be changed for the field functions which do not satisfy these conditions. In view of the fact that the conditions (3.1) make the action invariant with respect to the infinitesimal gauge trasformation, we now imagine to replace the mass term $`_{WM}`$ in (3.6) with a gauge invariant mass term which is equal to $`_{WM}`$ when the conditions (3.1) are satisfied. Thus, analogous to the case in the FadeevโPopov method \[1,3,11-16\], $`\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]`$ should be gauge invariant and make the following equation valid for an arbitrary gauge invariant quantity $`๐ช(๐ฒ,\overline{\psi },\psi )`$
$`{\displaystyle ๐[๐ฒ,\overline{\psi },\psi ]\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]\underset{a^{},x^{}}{}\delta \left(\mathrm{\Phi }_a^{}(x^{})\right)\delta \left(\mathrm{\Phi }_y(x^{})\right)๐ช(๐ฒ,\overline{\psi },\psi )\mathrm{exp}\{\mathrm{i}\stackrel{~}{I}\}}`$
$`{\displaystyle ๐[๐ฒ,\overline{\psi },\psi ]๐ช(๐ฒ,\overline{\psi },\psi )\mathrm{exp}\{\mathrm{i}\stackrel{~}{I}\}},`$
where $`\stackrel{~}{I}`$ is a gauge invariant action constructed by replacing $`_{WM}`$ with the imagined mass term. This means that the weight factor $`\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]`$ can be determined according to the FadeevโPopov equation of the following form
$`\mathrm{\Delta }[๐ฒ,\overline{\psi },\psi ]{\displaystyle \underset{z}{}d\mathrm{\Omega }(z)\underset{\sigma ,x}{}\delta \left(\mathrm{\Phi }_\sigma ^\mathrm{\Omega }(x)\right)}=1.`$ (3.7)
where $`\sigma `$ stands for $`1,2,3,y`$, $`\mathrm{\Phi }_\sigma ^\mathrm{\Omega }(x)`$ is the result of acting on $`\mathrm{\Phi }_\sigma (x)`$ with a gauge transformation having the parameters of the element $`\mathrm{\Omega }(x)`$ of the gauge group, $`d\mathrm{\Omega }(z)`$ is the volume element of the group integral. It follows that with the FโP ghost fields $`C_a(x)`$, $`C_y(x)`$, $`\overline{C}_a(x)`$, $`\overline{C}_y(x)`$ as new variables, one can express the ghost Lagrangian as
$`^{(C)}(x)=\overline{C}_a(x)\mathrm{\Delta }\mathrm{\Phi }_a(x)+\overline{C}_y(x)\mathrm{\Delta }\mathrm{\Phi }_y(x),`$ (3.8)
where $`\mathrm{\Delta }\mathrm{\Phi }_a(x)`$, $`\mathrm{\Delta }\mathrm{\Phi }_y(x)`$ are defined by the BRST transformtion of $`\mathrm{\Phi }_a(x)`$ and $`\mathrm{\Phi }_y(x)`$ so that
$`\delta _B\mathrm{\Phi }_a(x)=\delta \zeta \mathrm{\Delta }\mathrm{\Phi }_a(x),\delta _B\mathrm{\Phi }_y(x)=\delta \zeta \mathrm{\Delta }\mathrm{\Phi }_y(x),`$ (3.9)
where $`\delta \zeta `$ is an infinitesimal fermionic parameter independent of $`x`$. The BRST transformation of the gauge fields or matter fields is nothing but the infinitesimal gauge transformation with $`\delta \theta _a`$ and $`\delta \theta _y`$ equal to $`g\delta \zeta C_a`$ and $`g_1\delta \zeta C_y`$ respectively. Namely
$`\delta _BW_a^\mu (x)=\delta \zeta \mathrm{\Delta }W_a^\mu (x)=\delta \zeta D_{ab}^\mu C_b(x),`$ (3.10)
$`\delta _BW_y^\mu (x)=\delta \zeta \mathrm{\Delta }W_y^\mu (x)=\delta \zeta ^\mu C_y(x),`$ (3.11)
$`\delta _B\psi (x)=\delta \zeta \mathrm{\Delta }\psi (x),\delta _B\overline{\psi }(x)=\delta \zeta \mathrm{\Delta }\overline{\psi }(x),`$ (3.12)
where
$`D_{ab}^\mu (x)=\delta _{ab}^\mu +gf_{abc}A_c^\mu (x),`$
$`\mathrm{\Delta }\nu _L(x)={\displaystyle \frac{i}{2}}gC_1(x)e_L(x){\displaystyle \frac{1}{2}}gC_2(x)e_L(x){\displaystyle \frac{i}{2}}gC_3(x)\nu _L(x)+{\displaystyle \frac{i}{2}}g_1C_y(x)\nu _L(x),`$
$`\mathrm{\Delta }e_L(x)={\displaystyle \frac{i}{2}}gC_1(x)\nu _L(x)+{\displaystyle \frac{1}{2}}gC_2(x)\nu _L(x)+{\displaystyle \frac{i}{2}}gC_3(x)e_L(x)+{\displaystyle \frac{i}{2}}g_1C_y(x)e_L(x),`$
$`\mathrm{\Delta }e_R(x)=ig_1C_y(x)e_R(x),`$
$`\mathrm{\Delta }\overline{\nu }_L(x)={\displaystyle \frac{i}{2}}gC_1(x)\overline{e}_L(x){\displaystyle \frac{1}{2}}gC_2(x)\overline{e}_L(x)+{\displaystyle \frac{i}{2}}gC_3(x)\overline{\nu }_L(x){\displaystyle \frac{i}{2}}g_1C_y(x)\overline{\nu }_L(x),`$
$`\mathrm{\Delta }\overline{e}_L(x)={\displaystyle \frac{i}{2}}gC_1(x)\overline{\nu }_L(x)+{\displaystyle \frac{1}{2}}gC_2(x)\overline{\nu }_L(x){\displaystyle \frac{i}{2}}gC_3(x)\overline{e}_L(x){\displaystyle \frac{i}{2}}g_1C_y(x)\overline{e}_L(x),`$
$`\mathrm{\Delta }\overline{e}_R(x)=ig_1C_y(x)\overline{e}_R(x).`$
$`C_a(x)`$ and $`C_y(x)`$ are also transformed as usual
$`\delta _BC_a(x)=\delta \zeta \mathrm{\Delta }C_a(x)=\delta \zeta {\displaystyle \frac{g}{2}}C_{abc}C_b(x)C_c(x),`$
$`\delta _BC_y(x)=0.`$
Now we can write $`\mathrm{\Delta }\mathrm{\Phi }_a(x)`$, $`\mathrm{\Delta }\mathrm{\Phi }_y(x)`$ as
$`\mathrm{\Delta }\mathrm{\Phi }_1=_\mu \mathrm{\Delta }W_1^\mu (x)+g_1\mathrm{\Delta }W_2^\mu (x)W_{y\mu }(x)+g_1W_{2\mu }(x)\mathrm{\Delta }W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}\mathrm{\Delta }f_1(x),`$ (3.13)
$`\mathrm{\Delta }\mathrm{\Phi }_2=_\mu \mathrm{\Delta }W_2^\mu (x)g_1\mathrm{\Delta }W_1^\mu (x)W_{y\mu }(x)g_1W_{1\mu }(x)\mathrm{\Delta }W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}\mathrm{\Delta }f_2(x),`$ (3.14)
$`\mathrm{\Delta }\mathrm{\Phi }_3=_\mu \mathrm{\Delta }W_3^\mu (x)+g_1\mathrm{\Delta }W_3^\mu (x)W_{y\mu }(x)+g_1W_{3\mu }(x)\mathrm{\Delta }W_y^\mu (x)+{\displaystyle \frac{g}{M^2}}\mathrm{\Delta }f_3(x),`$ (3.15)
$`\mathrm{\Delta }\mathrm{\Phi }_y=_\mu \mathrm{\Delta }W_y^\mu (x)+g\mathrm{\Delta }W_3^\mu (x)W_{y\mu }(x)+gW_{3\mu }(x)\mathrm{\Delta }W_y^\mu (x),`$ (3.16)
where
$`\mathrm{\Delta }f_1(x)={\displaystyle \frac{i}{2}}m\left\{\left(\mathrm{\Delta }\overline{\nu }_L(x)\right)e_R(x)\overline{\nu }_L(x)\mathrm{\Delta }e_R(x)\left(\mathrm{\Delta }\overline{e}_R(x)\right)\nu _L(x)+\overline{e}_R(x)\mathrm{\Delta }\nu _L(x)\right\},`$
$`\mathrm{\Delta }f_2(x)={\displaystyle \frac{1}{2}}m\left\{\left(\mathrm{\Delta }\overline{\nu }_L(x)\right)e_R(x)\overline{\nu }_L(x)\mathrm{\Delta }e_R(x)+\left(\mathrm{\Delta }\overline{e}_R(x)\right)\nu _L(x)\overline{e}_R(x)\mathrm{\Delta }\nu _L(x)\right\},`$
$`\mathrm{\Delta }f_3(x)={\displaystyle \frac{i}{2}}m\left\{\left(\mathrm{\Delta }\overline{e}_R(x)\right)e_L(x)\overline{e}_R(x)\mathrm{\Delta }e_L(x)\left(\mathrm{\Delta }\overline{e}_L(x)\right)e_R(x)+\overline{e}_L(x)\mathrm{\Delta }e_R(x)\right\}.`$
Since $`\mathrm{\Delta }W_a^\mu `$, $`\mathrm{\Delta }W_y^\mu `$, $`\mathrm{\Delta }\psi (x)`$, $`\mathrm{\Delta }\overline{\psi }(x)`$ and $`\mathrm{\Delta }C_a(x)`$ are BRST invariant, it is easy to see that $`\mathrm{\Delta }\mathrm{\Phi }_a(x)`$ and $`\mathrm{\Delta }\mathrm{\Phi }_y(x)`$ are also BRST invariant.
One can further generalized the theory by regarding as new variables the Lagrange multipliers $`\lambda _a(x)`$ and $`\lambda _y(x)`$ associated with the constraint conditions. Thus the total effective Lagrangian and action consist of these Lagrange multipliers, ghosts and the original variables, namely
$`_{\mathrm{eff}}(x)=(x)+^{(C)}(x)+\lambda _a(x)\mathrm{\Phi }_a(x)+\lambda _y(x)\mathrm{\Phi }_y(x),`$ (3.17)
$`I_{\mathrm{eff}}={\displaystyle d^4x_{\mathrm{eff}}(x)}.`$ (3.18)
Correspondingly, the path integral of the generating functional for the Green functions is
$`๐ต[\overline{\eta },\eta ,\overline{\chi },\chi ,J,j]={\displaystyle \frac{1}{N_\lambda }}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C,\lambda ]\mathrm{exp}\left\{\mathrm{i}\left(I_{\mathrm{eff}}+I_s\right)\right\}},`$ (3.19)
where $`N_\lambda `$ is a constant, $`I_s`$ is the source term in the action. They are defined by
$`N_\lambda ={\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C,\lambda ]\mathrm{exp}\left\{\mathrm{i}I_{\mathrm{eff}}\right\}},`$
$`I_s={\displaystyle }d^4x\{\overline{\eta }(x)\psi (x)+\overline{\psi }(x)\eta (x)+\overline{\chi }_a(x)C_a(x)+\overline{C}_a(x)\chi _a(x)+\overline{\chi }_y(x)C_y(x)`$
$`+\overline{C}_y(x)\chi _y(x)+J_a^\mu (x)W_{a\mu }(x)+J_y^\mu (x)W_{y\mu }(x)+j_a(x)\lambda _a(x)+J_y(x)\lambda _y(x)\},`$ (3.20)
where $`\overline{\eta }(x),\eta (x)\mathrm{}`$ stand for the sources. In particular, $`j_a(x)`$, $`j_y(x)`$ are the sources of $`\lambda _a(x)`$, $`\lambda _y(x)`$, respectively.
We now check the BRST invariance of the effective action $`I_{eff}`$ defined by (3.17) and (3.18). With $`\overline{C}_a(x)`$, $`\overline{C}_y(x)`$ transforming as
$$\delta _B\overline{C}_a(x)=\delta \zeta \lambda _a(x),\delta _B\overline{C}_y(x)=\delta \zeta \lambda _y(x).$$
and noticing the invariance of $`\mathrm{\Delta }\mathrm{\Phi }_a,\mathrm{\Delta }\mathrm{\Phi }_y`$, one has
$$\delta _Bd^4x^{(C)}(x)=d^4x\left\{\lambda _a(x)\delta _B\mathrm{\Phi }_a(x)\lambda _y(x)\delta _B\mathrm{\Phi }_y(x)\right\}.$$
Therefore
$$\delta _BI_{\mathrm{eff}}=\delta _Bd^4x\left\{_{WM}+_{\psi m}\right\}+d^4x\left\{\left(\delta _B\lambda _a(x)\right)\mathrm{\Phi }_a(x)+\left(\delta _B\lambda _y(x)\right)\mathrm{\Phi }_y(x)\right\}.$$
From this and the expression of $`\delta _BI_{WM}`$, it can be shown that the effective action is invariant, when the transformation of $`\lambda _a(x)`$ and $`\lambda _y(x)`$ are defined as
$`\delta _B\lambda _1(x)=\delta \zeta M^2C_1(x),`$
$`\delta _B\lambda _2(x)=\delta \zeta M^2C_2(x),`$
$`\delta _B\lambda _3(x)=\delta \zeta M^2C_3(x)\delta \zeta {\displaystyle \frac{g_1}{g}}M^2C_y(x),`$
$`\delta _B\lambda _y(x)=\delta \zeta {\displaystyle \frac{g_1^2}{g^2}}M^2C_y(x)\delta \zeta {\displaystyle \frac{g_1}{g}}M^2C_3(x).`$
IV. Renormalisability
Following the notations of Ref. , let $`W_{a\mu }(x),W_{y\mu }(x)`$, $`C_a(x),C_y(x),\mathrm{}`$ stand for the renormalized field founctions, $`g,g_1`$ and $`M`$ be renormalized parameters. By introducing the source terms of the composite field functions $`\mathrm{\Delta }W_a^\mu `$, $`\mathrm{\Delta }W_y^\mu `$, $`\mathrm{\Delta }C_a(x)`$, $`\mathrm{\Delta }\psi (x)`$, $`\mathrm{\Delta }\overline{\psi }(x)`$ and the sources $`K_\mu ^a(x)`$, $`K_\mu ^y(x)`$, $`L_a(x)`$, $`n_\alpha (x)`$, $`l_\alpha (x)`$, $`p_\alpha (x)`$, $`n_\alpha ^{}(x)`$, $`l_\alpha ^{}(x)`$ and $`p_\alpha ^{}(x)`$, the effective Lagrangian without counterterm becomes
$`_{eff}^{[0]}(x)`$ $`=`$ $`\lambda _a(x)\mathrm{\Phi }_a(x)+\lambda _y(x)\mathrm{\Phi }_y(x)+_{WL}(x)+_{WY}(x)`$ (4.1)
$`+_{WM}(x)+^{(C)}(x)+_\psi (x)+_{\psi m}(x)+_{\psi W}(x)`$
$`+K_\mu ^a(x)\mathrm{\Delta }W_a^\mu (x)+K_\mu ^y(x)\mathrm{\Delta }W_y^\mu (x)+L_a(x)\mathrm{\Delta }C_a(x)`$
$`+n_\alpha (x)\mathrm{\Delta }\nu _{L\alpha }(x)+l_\alpha (x)\mathrm{\Delta }e_{L\alpha }(x)+p_\alpha (x)\mathrm{\Delta }e_{R\alpha }(x)`$
$`+n_\alpha ^{}(x)\mathrm{\Delta }\overline{\nu }_{L\alpha }(x)+l_\alpha ^{}(x)\mathrm{\Delta }\overline{e}_{L\alpha }(x)+p_\alpha ^{}(x)\mathrm{\Delta }\overline{e}_{R\alpha }(x).`$
The complete effective Lagrangian is the sum of $`_{eff}^{[0]}`$ and the counterterm $`_{count}`$
$`_{\mathrm{eff}}=_{\mathrm{eff}}^{[0]}+_{count}.`$ (4.2)
With (4.1), the generating functional for Green functions is defined as
$`๐ต^{[0]}[\overline{\eta },\eta ,\overline{\chi },\chi ,J,j,K,L,n,l,p,n^{},l^{},p^{}]={\displaystyle \frac{1}{N}}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C,\lambda ]\mathrm{exp}\left\{\mathrm{i}\left(I_{eff}^{[0]}+I_s\right)\right\}},`$ (4.3)
$`I_{eff}^{[0]}`$ is the effective action $`d^4x_{eff}^{[0]}(x)`$, $`N`$ is a constant to make $`๐ต^{[0]}`$ equal to $`1`$ in the absence of the sources, $`I_s`$ is the source term
$`I_s={\displaystyle }d^4x\{\overline{\eta }(x)\psi (x)+\overline{\psi }(x)\eta (x)+\overline{\chi }_a(x)C_a(x)+\overline{C}_a(x)\chi _a(x)+\overline{\chi }_y(x)C_y(x)`$
$`+\overline{C}_y(x)\chi _y(x)+J_a^\mu (x)W_{a\mu }(x)+J_y^\mu (x)W_{y\mu }(x)+j_a(x)\lambda _a(x)+j_y(x)\lambda _y(x)\},`$
where $`\overline{\eta }\psi `$ and $`\overline{\psi }\eta `$ stand for
$`\overline{\eta }\psi =\overline{\eta }_\alpha ^{(\nu )}\nu _{L\alpha }+\overline{\eta }_\alpha ^{(l)}e_{L\alpha }+\overline{\eta }_\alpha ^{(r)}e_{R\alpha },`$
$`\overline{\psi }\eta =\overline{\nu }_{L\alpha }\eta _\alpha ^{(\nu )}+\overline{e}_{L\alpha }\eta _\alpha ^{(l)}+\overline{e}_{R\alpha }\eta _\alpha ^{(r)}.`$
Denoting by $`๐ฒ^{[0]}`$ and $`\mathrm{\Gamma }^{[0]}`$ the generating functionals for connected Green functions and regular vertex functions respectively, one has
$`๐ต^{[0]}=\mathrm{exp}\left\{\mathrm{i}๐ฒ^{[0]}[\overline{\eta },\eta ,\overline{\chi },\chi ,J,j,K,L,n,l,p,n^{},l^{},p^{}]\right\},`$ (4.4)
$`\mathrm{\Gamma }^{[0]}[\stackrel{~}{\psi },\stackrel{~}{\overline{\psi }},\stackrel{~}{W},\stackrel{~}{\overline{C}},\stackrel{~}{C},\stackrel{~}{\lambda },K,L,n,l,p,n^{},l^{},p^{}]`$
$`=๐ฒ^{[0]}{\displaystyle }d^4x[J_a^\mu \stackrel{~}{W}_{a\mu }+J_y^\mu \stackrel{~}{W}_{y\mu }+j_a\stackrel{~}{\lambda }_a+j_y\stackrel{~}{\lambda }_y+\overline{\chi }_a\stackrel{~}{C}_a+\stackrel{~}{\overline{C}}_a\chi _a+\overline{\chi }_y\stackrel{~}{C}_y`$
$`+\stackrel{~}{\overline{C}}_y\chi _y+\overline{\eta }^{(\nu )}\stackrel{~}{\nu }_L+\overline{\eta }^{(l)}\stackrel{~}{e}_L+\overline{\eta }^{(r)}\stackrel{~}{e}_R+\stackrel{~}{\overline{\nu }}_L\eta ^{(\nu )}+\stackrel{~}{\overline{e}}_L\eta ^{(l)}+\stackrel{~}{\overline{e}}_R\eta ^{(r)}],`$ (4.5)
where $`\stackrel{~}{W}_{a\mu }`$, $`\stackrel{~}{\nu }_L`$, $`\mathrm{}`$ are the so-called classical fields defined by
$`\stackrel{~}{W}_{a\mu }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta J_a^\mu (x)}},\stackrel{~}{\lambda }_a(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta j_a(x)}},\stackrel{~}{C}_a(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \overline{\chi }_a(x)}},`$
$`\stackrel{~}{\overline{C}}_a(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \chi _a(x)}},\stackrel{~}{W}_{y\mu }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta J_y^\mu (x)}},\stackrel{~}{\lambda }_y={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta j_y(x)}},`$
$`\stackrel{~}{C}_y(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \overline{\chi }_y(x)}},\stackrel{~}{\overline{C}}_y(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \chi _y(x)}},\stackrel{~}{\nu }_{L\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \overline{\eta }_\alpha ^{(\nu )}(x)}},`$
$`\stackrel{~}{e}_{L\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \overline{\eta }_\alpha ^{(l)}(x)}},\stackrel{~}{e}_{R\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \overline{\eta }_\alpha ^{(r)}(x)}},\stackrel{~}{\overline{\nu }}_{L\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \eta _\alpha ^{(\nu )}(x)}},`$
$`\stackrel{~}{\overline{e}}_{L\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \eta _\alpha ^{(l)}(x)}},\stackrel{~}{\overline{e}}_{R\alpha }(x)={\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta \eta _\alpha ^{(r)}(x)}},`$
Therefore
$`J_a^\mu (x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{W}_{a\mu }(x)}},j_a(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}},\overline{\chi }_a(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{C}_a(x)}},`$
$`\chi _a(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_a(x)}},J_y^\mu (x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{W}_{y\mu }(x)}},j_y(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_y(x)}},`$
$`\overline{\chi }_y(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{C}_y(x)}},\chi _y(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_y(x)}},\eta _\alpha ^{(\nu )}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{\nu }}_{L\alpha }(x)}},`$
$`\eta _\alpha ^{(l)}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{e}}_{L\alpha }(x)}},\eta _\alpha ^{(r)}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{e}}_{R\alpha }(x)}},\overline{\eta }_\alpha ^{(\nu )}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\nu }_{L\alpha }(x)}},`$
$`\overline{\eta }_\alpha ^{(l)}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{e}_{L\alpha }(x)}},\overline{\eta }_\alpha ^{(r)}(x)={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{e}_{R\alpha }(x)}}.`$
Besides, for $`K_\mu ^a,L_a`$ $`\mathrm{}`$, the spectators in the Legendre transtrormation, one has
$`{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta K_\mu ^a(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^a(x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta K_\mu ^y(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta L_a(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta L_a(x)}},`$
$`{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta n_\alpha (x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha (x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta l_\alpha (x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha (x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta p_\alpha (x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}},`$
$`{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta n_\alpha ^{}(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha ^{}(x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta l_\alpha ^{}(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha ^{}(x)}},{\displaystyle \frac{\delta ๐ฒ^{[0]}}{\delta p_\alpha ^{}(x)}}={\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}.`$
In order to find the SlavnovโTaylor identity satisfied by the generating functional for the regular vertex functions, we change the variables in the path integral of $`๐ต^{[0]}`$ as follows
$`W_a^\mu (x)W_a^\mu (x)+\delta \zeta \mathrm{\Delta }W_a^\mu (x),W_y^\mu (x)W_y^\mu (x)+\delta \zeta \mathrm{\Delta }W_y^\mu (x),`$
$`C_a(x)C_a(x)+\delta \zeta \mathrm{\Delta }C_a(x),C_y(x)C_y(x),`$
$`\overline{C}_a(x)\overline{C}_a(x)\delta \zeta \lambda _a(x),\overline{C}_y(x)\overline{C}_y(x)\delta \zeta \lambda _y(x),`$
$`\psi (x)\psi (x)+\delta \zeta \mathrm{\Delta }\psi (x),\overline{\psi }(x)\overline{\psi }(x)+\delta \zeta \mathrm{\Delta }\overline{\psi }(x),`$
$`\lambda _a(x)\lambda _a(x),\lambda _y(x)\lambda _y(x).`$
The the changes in $`I_s`$ and $`_{WM}`$ lead to
$`{\displaystyle }d^4x\{{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^a(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{W}_a^\mu (x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{W}_y^\mu (x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta L_a(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{C}_a(x)}}`$
$`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\nu }_{L\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha (x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{e}_{L\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha (x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{e}_{R\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}}`$
$`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{\nu }}_{L\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{e}}_{L\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{e}}_{R\alpha }(x)}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}`$
$`\stackrel{~}{\lambda }_a(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_a(x)}}\stackrel{~}{\lambda }_y(x)+{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_y(x)}}\mathrm{\Delta }_{WM}(x)^{[0]}\mathrm{\Delta }_{\psi m}(x)^{[0]}\}=0,`$ (4.6)
where
$`\mathrm{\Delta }_{WM}(x)^{[0]}={\displaystyle \frac{1}{N๐ต^{[0]}}}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C]\mathrm{\Delta }_{WM}(x)\mathrm{exp}\left\{\mathrm{i}\left(I_{\mathrm{eff}}^{[0]}+I_s\right)\right\}},`$
$`\mathrm{\Delta }_{WM}(x)^{[0]}={\displaystyle \frac{1}{N๐ต^{[0]}}}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C]\mathrm{\Delta }_{\psi m}(x)\mathrm{exp}\left\{\mathrm{i}\left(I_{\mathrm{eff}}^{[0]}+I_s\right)\right\}}.`$
With the definitions of $`\mathrm{\Delta }_{WM}(x)`$ and $`\mathrm{\Delta }_{WM}(x)`$
$`\delta _B_{WM}(x)=\delta \zeta \mathrm{\Delta }_{WM}(x),\delta _B_{\psi m}(x)=\delta \zeta \mathrm{\Delta }_{\psi m}(x),`$
one can write
$`\mathrm{\Delta }_{WM}(x)^{[0]}`$ $`=`$ $`M^2\stackrel{~}{W}_{a\mu }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^a(x)}}+M^2\left({\displaystyle \frac{g_1}{g}}\right)^2\stackrel{~}{W}_{y\mu }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}`$
$`M^2{\displaystyle \frac{g_1}{g}}\stackrel{~}{W}_{y\mu }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^3(x)}}M^2{\displaystyle \frac{g_1}{g}}\stackrel{~}{W}_{3\mu }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}},`$
$`\mathrm{\Delta }_{\psi m}(x)^{[0]}`$ $`=`$ $`m{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha ^{}(x)}}\stackrel{~}{e}_{R\alpha }(x)+m\stackrel{~}{\overline{e}}_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}}`$
$`m{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}\stackrel{~}{e}_{L\alpha }(x)+m\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha (x)}}.`$
Next, from the invariance of the path integral of $`๐ต^{[0]}`$ with respect to the translation of the integration variables $`\overline{C}_a(x)`$, $`\overline{C}_y(x)`$, $`\lambda _a(x)`$ and $`\lambda _y(x)`$, one can get a set of auxiliary identities
$`{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_1(x)}}_\mu {\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^1(x)}}g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^2(x)}}g_1\stackrel{~}{W}_{2\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{\nu }}_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}}+\stackrel{~}{\nu }_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha (x)}}\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha ^{}(x)}}\right\}=0,`$ (4.7)
$`{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_2(x)}}_\mu {\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^2(x)}}+g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^1(x)}}+g_1\stackrel{~}{W}_{1\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{\nu }}_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}}\stackrel{~}{\nu }_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}+\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha (x)}}\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta n_\alpha ^{}(x)}}\right\}=0,`$ (4.8)
$`{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_3(x)}}_\mu {\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^3(x)}}g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^3(x)}}g_1\stackrel{~}{W}_{3\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha (x)}}+\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta l_\alpha ^{}(x)}}\stackrel{~}{\overline{e}}_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha (x)}}\stackrel{~}{e}_{L\alpha }(x){\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta p_\alpha ^{}(x)}}\right\}=0,`$ (4.9)
$`{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\overline{C}}_y(x)}}_\mu {\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}g\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^3(x)}}g\stackrel{~}{W}_{3\mu }{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta K_\mu ^y(x)}}=0,`$ (4.10)
and
$`{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}}=\mathrm{\Phi }_a(x)^{[0]},{\displaystyle \frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_y(x)}}=\mathrm{\Phi }_y(x)^{[0]}.`$ (4.11)
where
$`\mathrm{\Phi }_a(x)^{[0]}={\displaystyle \frac{1}{N๐ต^{[0]}}}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C,\lambda ]\mathrm{\Phi }_a(x)\mathrm{exp}\left\{\mathrm{i}\left(I_{\mathrm{eff}}^{[0]}+I_s\right)\right\}},`$ (4.12)
$`\mathrm{\Phi }_y(x)^{[0]}={\displaystyle \frac{1}{N๐ต^{[0]}}}{\displaystyle ๐[\overline{\psi },\psi ,๐ฒ,\overline{C},C,\lambda ]\mathrm{\Phi }_y(x)\mathrm{exp}\left\{\mathrm{i}\left(I_{\mathrm{eff}}^{[0]}+I_s\right)\right\}}.`$ (4.13)
Let $`\stackrel{~}{\mathrm{\Phi }}_a(x)`$, $`\stackrel{~}{\mathrm{\Phi }}_y(x)`$, $`\stackrel{~}{}_{WM}`$ and $`\stackrel{~}{}_{\psi m}`$ be the results obtained from $`\mathrm{\Phi }_a(x)`$, $`\mathrm{\Phi }_y(x)`$, $`_{WM}`$ and $`_{\psi m}`$ by replacing the field functions with the classical field functions and define
$`\overline{\mathrm{\Gamma }}^{[0]}=\mathrm{\Gamma }^{[0]}{\displaystyle d^4x\left\{\stackrel{~}{\lambda }_a(x)\stackrel{~}{\mathrm{\Phi }}_a(x)+\stackrel{~}{\lambda }_y(x)\stackrel{~}{\mathrm{\Phi }}_y(x)+\stackrel{~}{}_{WM}+\stackrel{~}{}_{\psi m}\right\}},`$ (4.14)
Thus, from (4.6)โ(4.11), one gets
$`{\displaystyle }d^4x\{{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^a(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{W}_a^\mu (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{W}_y^\mu (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta L_a(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{C}_a(x)}}`$
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\nu }_{L\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{e}_{L\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta l_\alpha (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{e}_{R\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha (x)}}`$
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{\nu }}_{L\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{e}}_{L\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta l_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{e}}_{R\alpha }(x)}}{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha ^{}(x)}}\}=0.`$ (4.15)
and
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}}=\mathrm{\Phi }_a(x)^{[0]}\stackrel{~}{\mathrm{\Phi }}_a(x),{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\lambda }_y(x)}}=\mathrm{\Phi }_y(x)^{[0]}\stackrel{~}{\mathrm{\Phi }}_y(x),`$ (4.16)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{C}}_1(x)}}_\mu {\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^1(x)}}g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^2(x)}}g_1\stackrel{~}{W}_{2\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{\nu }}_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha (x)}}+\stackrel{~}{\nu }_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha ^{}(x)}}\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha (x)}}\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha ^{}(x)}}\right\}=0,`$ (4.17)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{C}}_2(x)}}_\mu {\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^2(x)}}+g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^1(x)}}+g_1\stackrel{~}{W}_{1\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{\nu }}_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha (x)}}\stackrel{~}{\nu }_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha ^{}(x)}}+\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha (x)}}\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta n_\alpha ^{}(x)}}\right\}=0,`$ (4.18)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{C}}_3(x)}}_\mu {\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^3(x)}}g_1\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^3(x)}}g_1\stackrel{~}{W}_{3\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}`$
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\stackrel{~}{\overline{e}}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta l_\alpha (x)}}+\stackrel{~}{e}_{R\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta l_\alpha ^{}(x)}}\stackrel{~}{\overline{e}}_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha (x)}}\stackrel{~}{e}_{L\alpha }(x){\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta p_\alpha ^{}(x)}}\right\}=0,`$ (4.19)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\overline{C}}_y(x)}}_\mu {\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}g\stackrel{~}{W}_{y\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^3(x)}}g\stackrel{~}{W}_{3\mu }{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta K_\mu ^y(x)}}=0.`$ (4.20)
As mentioned earlier our intention to use the generalized form of the theory containing $`\lambda _a`$, $`\lambda _y`$ and their sources is to study the Renormalisability of the theory for which such sources are absent from the generating functional for the Green functions and therefore $`\mathrm{\Phi }_a(x)^{[0]}`$ and $`\mathrm{\Phi }_y(x)^{[0]}`$ are equal to zero. We now, according to $`(4.11)`$, let vanish $`\frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}`$ and $`\frac{\delta \mathrm{\Gamma }^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}`$ to make $`\mathrm{\Phi }_a(x)^{[0]}`$ and $`\mathrm{\Phi }_y(x)^{[0]}`$ equal to zero. This means
$`\stackrel{~}{\mathrm{\Phi }}_a(x)=0,\stackrel{~}{\mathrm{\Phi }}_y(x)=0,`$ (4.21)
and
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\lambda }_a(x)}}=0,{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \stackrel{~}{\lambda }_y(x)}}=0.`$ (4.22)
In the following we will denote by $`\overline{\mathrm{\Gamma }}^{[0]}[\psi ,\overline{\psi },W,\overline{C},C,\lambda ,K,L,n,l,p,n^{},l^{},p^{}]`$ the functional that is obtained from $`\overline{\mathrm{\Gamma }}^{[0]}[\stackrel{~}{\psi },\stackrel{~}{\overline{\psi }},\stackrel{~}{W},\stackrel{~}{\overline{C}},\stackrel{~}{C},\stackrel{~}{\lambda },K,\mathrm{}]`$ by replacing the classical field functions with the usual field functions. Assume that the dimensional regularization method is used and the SlavnovโTaylor identity and the auxiliary identities are guaranteed. Denote the tree part and one loop part of $`\overline{\mathrm{\Gamma }}^{[0]}`$ by $`\overline{\mathrm{\Gamma }}_0^{[0]}`$ and $`\overline{\mathrm{\Gamma }}_1^{[0]}`$ respectively. $`\overline{\mathrm{\Gamma }}_0^{[0]}`$ is thus the modified action $`\overline{I}_{eff}^{[0]}`$ obtained from $`I_{eff}^{[0]}`$ by excluding the mass term and $`(\lambda _a,\lambda _y)`$ terms. From $`(4.15)`$ and $`(4.17)(4.22)`$ one has
$`\mathrm{\Phi }_a(x)=0,\mathrm{\Phi }_y(x)=0,`$ (4.23)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \lambda _a(x)}}=0,{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[0]}}{\delta \lambda _y(x)}}=0,`$ (4.24)
$`\mathrm{\Lambda }_{op}\overline{\mathrm{\Gamma }}_0^{[0]}=0,`$
and
$`\overline{\mathrm{\Gamma }}_0^{[0]}\overline{\mathrm{\Gamma }}_1^{[0]}+\overline{\mathrm{\Gamma }}_1^{[0]}\overline{\mathrm{\Gamma }}_0^{[0]}=\mathrm{\Lambda }_{op}\overline{\mathrm{\Gamma }}_1^{[0]}=0,`$ (4.25)
$`\mathrm{\Sigma }_a(x)\overline{\mathrm{\Gamma }}^{[0]}=0,\mathrm{\Sigma }_y(x)\overline{\mathrm{\Gamma }}^{[0]}=0,`$ (4.26)
where $`\mathrm{\Lambda }_{op}`$,$`\mathrm{\Sigma }_a(x)`$ and $`\mathrm{\Sigma }_y(x)`$ are defined by
$`\mathrm{\Lambda }_{op}`$ $`=`$ $`{\displaystyle }d^4x\{{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta K_\mu ^a(x)}}{\displaystyle \frac{\delta }{\delta W_a^\mu (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta W_a^\mu (x)}}{\displaystyle \frac{\delta }{\delta K_\mu ^a(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta K_\mu ^y(x)}}{\displaystyle \frac{\delta }{\delta W_y^\mu (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta W_y^\mu (x)}}{\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}`$ (4.27)
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta L_a(x)}}{\displaystyle \frac{\delta }{\delta C_a(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_a(x)}}{\displaystyle \frac{\delta }{\delta L_a(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \nu _{L\alpha }(x)}}{\displaystyle \frac{\delta }{\delta n_\alpha (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta n_\alpha (x)}}{\displaystyle \frac{\delta }{\delta \nu _{L\alpha }(x)}}`$
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta e_{L\alpha }(x)}}{\displaystyle \frac{\delta }{\delta l_\alpha (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta l_\alpha (x)}}{\displaystyle \frac{\delta }{\delta e_{L\alpha }(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta e_{R\alpha }(x)}}{\displaystyle \frac{\delta }{\delta p_\alpha (x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta p_\alpha (x)}}{\displaystyle \frac{\delta }{\delta e_{R\alpha }(x)}}`$
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{\nu }_{L\alpha }(x)}}{\displaystyle \frac{\delta }{\delta n_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta n_\alpha ^{}(x)}}{\displaystyle \frac{\delta }{\delta \overline{\nu }_{L\alpha }(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{e}_{L\alpha }(x)}}{\displaystyle \frac{\delta }{\delta l_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta l_\alpha ^{}(x)}}{\displaystyle \frac{\delta }{\delta \overline{e}_{L\alpha }(x)}}`$
$`+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{e}_{R\alpha }(x)}}{\displaystyle \frac{\delta }{\delta p_\alpha ^{}(x)}}+{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta p_\alpha ^{}(x)}}{\displaystyle \frac{\delta }{\delta \overline{e}_{R\alpha }(x)}}\},`$
$`\mathrm{\Sigma }_1(x)`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta \overline{C}_1(x)}}_\mu {\displaystyle \frac{\delta }{\delta K_\mu ^1(x)}}g_1W_{y\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^2(x)}}g_1W_{2\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}`$ (4.28)
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\overline{\nu }_{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha (x)}}+\nu _{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha ^{}(x)}}\overline{e}_{R\alpha }(x){\displaystyle \frac{\delta }{\delta n_\alpha (x)}}e_{R\alpha }(x){\displaystyle \frac{\delta }{\delta n_\alpha ^{}(x)}}\right\}=0,`$
$`\mathrm{\Sigma }_2(x)`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta \overline{C}_2(x)}}_\mu {\displaystyle \frac{\delta }{\delta K_\mu ^2(x)}}+g_1W_{y\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^1(x)}}+g_1W_{1\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}`$ (4.29)
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\overline{\nu }_{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha (x)}}\nu _{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha ^{}(x)}}+\overline{e}_{R\alpha }(x){\displaystyle \frac{\delta }{\delta n_\alpha (x)}}e_{R\alpha }(x){\displaystyle \frac{\delta }{\delta n_\alpha ^{}(x)}}\right\}=0,`$
$`\mathrm{\Sigma }_3(x)`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta \overline{C}_3(x)}}_\mu {\displaystyle \frac{\delta }{\delta K_\mu ^3(x)}}g_1W_{y\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^3(x)}}g_1W_{3\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}`$ (4.30)
$`+{\displaystyle \frac{i}{2}}{\displaystyle \frac{mg}{M^2}}\left\{\overline{e}_{R\alpha }(x){\displaystyle \frac{\delta }{\delta l_\alpha (x)}}+e_{R\alpha }(x){\displaystyle \frac{\delta }{\delta l_\alpha ^{}(x)}}\overline{e}_{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha (x)}}e_{L\alpha }(x){\displaystyle \frac{\delta }{\delta p_\alpha ^{}(x)}}\right\}=0,`$
$`\mathrm{\Sigma }_y(x)`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta \overline{C}_y(x)}}_\mu {\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}gW_{y\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^3(x)}}gW_{3\mu }{\displaystyle \frac{\delta }{\delta K_\mu ^y(x)}}.`$ (4.31)
The meaning of the notation $`AB`$ is
$`AB={\displaystyle }d^4x\{{\displaystyle \frac{\delta A}{\delta K_\mu ^a(x)}}{\displaystyle \frac{\delta B}{\delta W_a^\mu (x)}}+{\displaystyle \frac{\delta A}{\delta K_\mu ^y(x)}}{\displaystyle \frac{\delta B}{\delta W_y^\mu (x)}}+{\displaystyle \frac{\delta A}{\delta L_a(x)}}{\displaystyle \frac{\delta B}{\delta C_a(x)}}`$
$`+{\displaystyle \frac{\delta A}{\delta \nu _{L\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta n_\alpha (x)}}+{\displaystyle \frac{\delta A}{\delta e_{L\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta l_\alpha (x)}}+{\displaystyle \frac{\delta A}{\delta e_{R\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta p_\alpha (x)}}`$
$`+{\displaystyle \frac{\delta A}{\delta \overline{\nu }_{L\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta n_\alpha ^{}(x)}}+{\displaystyle \frac{\delta A}{\delta \overline{e}_{L\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta l_\alpha ^{}(x)}}+{\displaystyle \frac{\delta A}{\delta \overline{e}_{R\alpha }(x)}}{\displaystyle \frac{\delta B}{\delta p_\alpha ^{}(x)}}\}.`$ (4.32)
$`(4.24)(4.26)`$ are of course satisfied by the finite part and the pole part of $`\overline{\mathrm{\Gamma }}_1^{[0]}`$. Thus the equations of the pole part $`\overline{\mathrm{\Gamma }}_1^{[0]}`$ are
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_1^{[0]}}{\delta \lambda _a(x)}}=0,{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_1^{[0]}}{\delta \lambda _y(x)}}=0,`$ (4.33)
$`\mathrm{\Lambda }_{op}\overline{\mathrm{\Gamma }}_1^{[0]}=0,`$ (4.34)
$`\mathrm{\Sigma }_a(x)\overline{\mathrm{\Gamma }}_1^{[0]}=0,\mathrm{\Sigma }_y(x)\overline{\mathrm{\Gamma }}_1^{[0]}=0.`$ (4.35)
It is known that when $`m=0`$ the theory is renormalisable and $`\overline{\mathrm{\Gamma }}_{1,div}^{[0]}`$ is a combination of 5 independent terms. Now one can also find the corresponding solutions of equations $`(4.33)(4.35)`$. These solutions are as follows
$`T_{(1)}=T_{WL}T_{GL}T_{CK}+2I_m^{(C)},`$ (4.36)
$`T_{(2)}=T_{WY}T_{GY}T_{CKY},`$ (4.37)
$`T_{(3)}=T_{CK}+T_{CKY}+T_{nn^{}}+T_{ll^{}}+T_{pp^{}},`$ (4.38)
$`T_{(4)}=T_{\nu L}+T_{eL}T_{nn^{}}T_{ll^{}}I_m^{(C)},`$ (4.39)
$`T_{(5)}=T_{eR}T_{pp^{}}I_m^{(C)},`$ (4.40)
where
$`T_{GL}=g{\displaystyle \frac{\overline{\mathrm{\Gamma }}_0^{[0]}}{g}},T_{GY}=g_1{\displaystyle \frac{\overline{\mathrm{\Gamma }}_0^{[0]}}{g_1}},I_m^{(C)}=m{\displaystyle \frac{\overline{\mathrm{\Gamma }}_0^{[0]}}{m}}=u{\displaystyle \frac{\overline{\mathrm{\Gamma }}_0^{[0]}}{u}},`$
$`T_{WL}={\displaystyle d^4x\left\{W_a^\mu (x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta W_a^\mu (x)}+L_a(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta L_a(x)}\right\}},`$
$`T_{WY}={\displaystyle d^4xW_y^\mu (x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta W_y^\mu (x)}},`$
$`T_{CK}={\displaystyle d^4x\left\{\overline{C}_a(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{C}_a(x)}+C_a(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_a(x)}+K_\mu ^a(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta K_\mu ^a(x)}\right\}},`$
$`T_{CKY}={\displaystyle d^4x\left\{\overline{C}_y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{C}_(x)}+C_y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_y(x)}+K_\mu ^y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta K_\mu ^y(x)}\right\}},`$
$`T_{\nu L}={\displaystyle d^4x\left\{\nu _{L\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \nu _{L\alpha }(x)}+\overline{\nu }_{L\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{\nu }_{L\alpha }(x)}\right\}},`$
$`T_{eL}={\displaystyle d^4x\left\{e_{L\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta e_{L\alpha }(x)}+\overline{e}_{L\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{e}_{L\alpha }(x)}\right\}},`$
$`T_{eR}={\displaystyle d^4x\left\{e_{R\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta e_{R\alpha }(x)}+\overline{e}_{R\alpha }(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta \overline{e}_{R\alpha }(x)}\right\}},`$
$`T_{nn^{}}={\displaystyle d^4x\left\{n_\alpha (x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta n_\alpha (x)}+n_\alpha ^{}(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta n_\alpha ^{}(x)}\right\}},`$
$`T_{ll^{}}={\displaystyle d^4x\left\{l_\alpha (x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta l_\alpha (x)}+l_\alpha ^{}(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta l_\alpha ^{}(x)}\right\}},`$
$`T_{pp^{}}={\displaystyle d^4x\left\{p_\alpha (x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta p_\alpha (x)}+p_\alpha ^{}(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta p_\alpha ^{}(x)}\right\}},`$
where the parameter $`u`$ appearing in the expression of $`I_m^{(C)}`$ stands for $`m/M^2`$. Similar to the case of Ref. , $`T_{(3)}`$ is $`2\left(\overline{\mathrm{\Gamma }}_0^{[0]}I_{WL}I_{WY}I_\psi I_{\psi W}\right)`$. $`T_{(1)}`$ is a combination of $`I_{WL}`$, $`T_{(3)}`$ and $`d^4xC_y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_y(x)}`$. $`T_{(2)}`$ is a combination of $`I_{WY}`$ and $`d^4xC_y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_y(x)}`$. The sum of $`T_{(4)}`$ and $`T_{(5)}`$ is $`2\left(I_\psi +I_{\psi W}\right)`$. $`d^4xC_y(x)\frac{\delta \overline{\mathrm{\Gamma }}_0^{[0]}}{\delta C_y(x)}`$ and $`T_{(5)}`$ can be easily checked to satisfy $`(4.33)(4.35)`$.
Since $`(4.36)`$$`(4.40)`$ become the whole independent terms of $`\overline{\mathrm{\Gamma }}_{1,div}^{[0]}`$ when $`m=0`$, a new term that appears when $`m0`$ should include $`m`$ $`\overline{\psi }_\alpha `$ $`\psi _\alpha ^{}`$ as a factor and also satisfies $`(4.33)(4.35)`$. First, it is clearly not possible to form such a solution of $`(4.33)(4.35)`$ if negative dimension coefficients are excluded. Next, if $`M^2`$ is included as a factor in the definition of the ghost action so that the ghost fields become dimensionless, then the modified effective action (without $`\lambda `$ terms) does not contain negative dimension coefficients. Thus a new term that can appear in $`\overline{\mathrm{\Gamma }}_1^{[0]}`$ must be formed with some powers of $`\overline{C_\sigma }`$ $`C_\sigma ^{}`$ and a factor from $`m\overline{\psi }_\alpha `$$`\psi _\alpha ^{}`$, where $`\sigma `$ and $`\sigma ^{}`$ stand for $`1,2,3,y`$. However, $`(4.33)(4.35)`$ does not have such a solution neither. This can easily be seen from $`(4.35)`$. It follows that $`\overline{\mathrm{\Gamma }}_{1,div}^{[0]}`$ is a combination of (4.36)โ(4.40). Namely
$`\overline{\mathrm{\Gamma }}_{1,div}^{[0]}=\alpha _1^{(1)}T_{(1)}+\alpha _2^{(1)}T_{(2)}+\alpha _3^{(1)}T_{(3)}+\alpha _4^{(1)}T_{(4)}+\alpha _5^{(1)}T_{(5)},`$ (4.41)
where, $`\alpha _1^{(1)},\mathrm{},\alpha _5^{(1)}`$ are constants of order $`(\mathrm{})^1`$ and are divergent when the space-time dimension tends to $`4`$.
In order to cancel the one loop divergence the counterterm of order $`\mathrm{}^1`$ in the action should be chosen as
$`\delta I_{count}^{[1]}=\overline{\mathrm{\Gamma }}_{1,div}^{[0]}.`$ (4.42)
Since
$`\overline{I}_{eff}^{[0]}=\overline{\mathrm{\Gamma }}_0^{[0]},`$ (4.43)
it is known from $`(4.41)`$ that the sum of $`\overline{I}_{eff}^{[0]}`$ and $`\delta I_{count}^{[1]}`$, to order of $`\mathrm{}^1`$, can be written as
$`\overline{I}_{eff}^{[1]}[\psi ,\overline{\psi },W,C,\overline{C},K,L,n,l,p,n^{},l^{},p^{},g,g_1,u]`$
$`=\overline{I}_{\mathrm{eff}}^{[0]}[\psi ^{[0]},\overline{\psi }^{[0]},W^{[0]},C^{[0]},\overline{C}^{[0]},K^{[0]},L^{[0]},n^{[0]},n^{{}_{}{}^{}[0]},\mathrm{},g^{[0]},g_1^{[0]},u^{[0]}],`$ (4.44)
where the bare fields and the bare parameters (to order $`(\mathrm{})^1`$) are defined as
$`W_{a\mu }^{[0]}=(Z_3^{[1]})^{1/2}W_{a\mu }=\left(1\alpha _1^{(1)}\right)W_{a\mu },L_a^{[0]}=(Z_3^{[1]})^{1/2}L_a,`$ (4.45)
$`W_{y\mu }^{[0]}=(Z_3^{{}_{}{}^{}[1]})^{1/2}W_{y\mu }=\left(1\alpha _2^{(1)}\right)W_{y\mu },`$ (4.46)
$`C_a^{[0]}=(\stackrel{~}{Z}_3^{[1]})^{1/2}C_a=\left(1\alpha _3^{(1)}+\alpha _1^{(1)}\right)C_a,`$ (4.47)
$`\overline{C}_a^{[0]}=(\stackrel{~}{Z}_3^{[1]})^{1/2}\overline{C}_a,K_\mu ^{a[0]}=(\stackrel{~}{Z}_3^{[1]})^{1/2}K_\mu ^a,`$ (4.48)
$`C_y^{[0]}=(\stackrel{~}{Z}_3^{{}_{}{}^{}[1]})^{1/2}C_y=\left(1\alpha _3^{(1)}+\alpha _2^{(1)}\right)C_y,`$ (4.49)
$`\overline{C}_y^{[0]}=(\stackrel{~}{Z}_3^{{}_{}{}^{}[1]})^{1/2}\overline{C}_y,K_\mu ^{y[0]}=(\stackrel{~}{Z}_3^{[1]})^{1/2}K_\mu ^y,`$ (4.50)
$`\nu _L^{[0]}=(Z_{\nu L}^{[1]})^{1/2}\nu _L=\left(1\alpha _4^{(1)}\right)\nu _L,\overline{\nu }_L^{[0]}=(Z_{\nu L}^{[1]})^{1/2}\overline{\nu }_L,`$ (4.51)
$`e_L^{[0]}=(Z_{eL}^{[1]})^{1/2}e_L=(Z_{\nu L}^{[1]})^{1/2}e_L,\overline{e}_L^{[0]}=(Z_{eL}^{[1]})^{1/2}\overline{e}_L,`$ (4.52)
$`e_R^{[0]}=(Z_{eR}^{[1]})^{1/2}e_R=\left(1\alpha _5^{(1)}\right)e_R,\overline{e}_R^{[0]}=(Z_{eR}^{[1]})^{1/2}\overline{e}_R,`$ (4.53)
$`n^{[0]}=(Z_{(n)}^{[1]})^{1/2}n=\left(1\alpha _3^{(1)}+\alpha _4^{(1)}\right)n,n^{{}_{}{}^{}[0]}=(Z_{(n)}^{[1]})^{1/2}n^{},`$ (4.54)
$`l^{[0]}=(Z_{(l)}^{[1]})^{1/2}l=(Z_{(n)}^{[1]})^{1/2}l,l^{{}_{}{}^{}[0]}=(Z_{(l)}^{[1]})^{1/2}l^{},`$ (4.55)
$`p^{[0]}=(Z_{(p)}^{[1]})^{1/2}p=\left(1\alpha _3^{(1)}+\alpha _5^{(1)}\right)p,p^{{}_{}{}^{}[0]}=(Z_{(p)}^{[1]})^{1/2}p^{},`$ (4.56)
$`g^{[0]}=Z_g^{[1]}g=(Z_3^{[1]})^{1/2}g,g_1^{[0]}=Z_g^{{}_{}{}^{}[1]}g_1=(Z_3^{{}_{}{}^{}[1]})^{1/2}g_1,`$ (4.57)
$`u^{[0]}=\left(12\alpha _1^{(1)}+\alpha _4^{(1)}+\alpha _5^{(1)}\right)u.`$ (4.58)
Next let $`\mathrm{\Phi }_a^{[0]}`$ and $`\mathrm{\Phi }_y^{[0]}`$ be obtained from $`\mathrm{\Phi }_a`$ and $`\mathrm{\Phi }_y`$ by replacing the field functions and parameters with the bare field functions and bare parameters. From $`(4.45),(4.46)`$ and $`(4.57)`$ one has
$`\mathrm{\Phi }_a^{[0]}=(Z_3^{[1]})^{1/2}\mathrm{\Phi }_a,\mathrm{\Phi }_y^{[0]}=(Z_3^{{}_{}{}^{}[1]})^{1/2}\mathrm{\Phi }_y.`$ (4.59)
Thus by adding the mass terms and the $`\lambda `$ terms into $`\overline{I}_{eff}^{[1]}`$ and forming
$`I_{eff}^{[1]}=\overline{I}_{eff}^{[1]}+I_{WM}+I_{\psi m}+{\displaystyle d^4x\left\{\lambda _a(x)\mathrm{\Phi }_a(x)+\lambda _y(x)\mathrm{\Phi }_y(x)\right\}},`$ (4.60)
one gets
$`I_{eff}^{[1]}[\psi ,\overline{\psi },W,C,\overline{C},\lambda ,K,L,n,l,p,n^{},l^{},p^{},g,g_1,M,m]`$
$`=I_{eff}^{[0]}[\psi ^{[0]},\overline{\psi }^{[0]},W^{[0]},C^{[0]},\overline{C}^{[0]},\lambda ^{[0]},K^{[0]},L^{[0]},n^{[0]},n^{{}_{}{}^{}[0]},\mathrm{},g^{[0]},g_1^{[0]},M^{[0]},m^{[0]}],`$ (4.61)
where
$`M^{[0]}=(Z_3^{[1]})^{1/2}M,m^{[0]}=(Z_{eL}^{[1]})^{1/2}(Z_{eR}^{[1]})^{1/2}m,`$ (4.62)
and
$`\lambda _a^{[0]}=(Z_3^{[1]})^{1/2}\lambda _a,\lambda _y^{[0]}=(Z_3^{{}_{}{}^{}[1]})^{1/2}\lambda _y.`$ (4.63)
Obviously, if the action $`I_{eff}^{[1]}`$ is used to replace $`I_{eff}^{[0]}`$ in $`(4.3)`$ and define $`๐ต^{[1]}`$, $`\mathrm{\Gamma }^{[1]}`$ as well as
$`\overline{\mathrm{\Gamma }}^{[1]}=\mathrm{\Gamma }^{[1]}I_{WM}I_{\psi m}{\displaystyle d^4x\left\{\lambda _a(x)\mathrm{\Phi }_a(x)+\lambda _y(x)\mathrm{\Phi }_y(x)\right\}},`$ (4.64)
then one has
$`\overline{\mathrm{\Gamma }}^{[1]}[\psi ,\overline{\psi },W,C,\overline{C},\lambda ,K,L,n,l,p,n^{},l^{},p^{},g,g_1,u]`$
$`=\overline{\mathrm{\Gamma }}^{[0]}[\psi ^{[0]},\overline{\psi }^{[0]},W^{[0]},C^{[0]},\overline{C}^{[0]},\lambda ^{[0]},K^{[0]},L^{[0]},n^{[0]},n^{{}_{}{}^{}[0]},\mathrm{},g^{[0]},g_1^{[0]},u^{[0]}].`$ (4.65)
From this it is easy to check that, to order $`\mathrm{}^1`$, $`\overline{\mathrm{\Gamma }}^{[1]}`$ is finite. Moreover, by changing into bare fields and bare parameters the fields and parameters in $`(4.15)`$ $`(4.22)`$ and then transforming them back into the renormalized fields and renormalized parameters according to $`(4.45)`$$`(4.59)`$, one can see that, under condition $`(4.23)`$, $`\overline{\mathrm{\Gamma }}^{[1]}`$ also satisfies
$`\mathrm{\Lambda }_{op}\overline{\mathrm{\Gamma }}^{[1]}=0,`$ (4.66)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[1]}}{\delta \lambda _a(x)}}=0,{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}^{[1]}}{\delta \lambda _y(x)}}=0,`$ (4.67)
$`\mathrm{\Sigma }_a(x)\overline{\mathrm{\Gamma }}^{[1]}=0,\mathrm{\Sigma }_y(x)\overline{\mathrm{\Gamma }}^{[1]}=0.`$ (4.68)
We can now use the inductive method and follow the steps of Ref. to complete the proof of renormalisability. Assume that up to $`n`$ loop the theory has been proved to be renormalisable by introducing the counterterm
$$I_{\mathrm{count}}^{[n]}=\underset{l=1}{\overset{n}{}}\delta I_{\mathrm{count}}^{[l]},$$
where $`\delta I_{count}^{[l]}`$ is the counterterm of order $`\mathrm{}^l`$ and has the form of (4.41),(4.42). Therefore the modified generating functional $`\overline{\mathrm{\Gamma }}^{[n]}`$ for the regular vertex, defined by the action
$$I_{\mathrm{eff}}^{[n]}=I_{\mathrm{eff}}^{[0]}+I_{\mathrm{count}}^{[n]},$$
satisfied equations $`(4.66)(4.68)`$ (under $`(4.23)`$) and, to order $`\mathrm{}^n`$, is finite. This also means that the fields or parameters in each of the following brackets have the same renormalization factor
$$(W_{a\mu }^{[0]},L_a),(C_a,\overline{C}_a,K_\mu ^a),(C_y,\overline{C}_y,K_\mu ^y),(\nu _L,\overline{\nu }_L,e_L,\overline{e}_L),(e_R,\overline{e}_R),(n,n^{},l,l^{}),(p,p^{}),(\lambda ,M,g),$$
and that
$`Z_g^{{}_{}{}^{}[n]}(Z_3^{{}_{}{}^{}[n]})^{1/2}=1,Z_g^{[n]}(Z_3^{[n]})^{1/2}=1,`$
$`Z_3^{[n]}\stackrel{~}{Z}_3^{[n]}=\stackrel{~}{Z}_3^{{}_{}{}^{}[n]}\stackrel{~}{Z}_3^{{}_{}{}^{}[n]}=Z_{\nu L}^{[n]}Z_{(n)}^{[n]}=Z_{eR}^{[n]}Z_{(p)}^{[n]}.`$
Denote by $`\overline{\mathrm{\Gamma }}_k^{[n]}`$ the part of order $`\mathrm{}^k`$ in $`\overline{\mathrm{\Gamma }}^{[n]}`$. For $`kn`$, $`\overline{\mathrm{\Gamma }}_k^{[n]}`$ is equal to $`\overline{\mathrm{\Gamma }}_k^{[k]}`$, because it can not contain the contribution of a counterterm of order $`\mathrm{}^{k+1}`$ or higher. Thus on expanding $`\overline{\mathrm{\Gamma }}^{[n]}`$ to order $`\mathrm{}^{n+1}`$ one has
$$\overline{\mathrm{\Gamma }}^{[n]}=\underset{k=0}{\overset{n}{}}\overline{\mathrm{\Gamma }}_k^{[k]}+\overline{\mathrm{\Gamma }}_{n+1}^{[n]}+\mathrm{}.$$
Using this and extracting the terms of order $`\mathrm{}^{(n+1)}`$ from the equations satisfied by $`\overline{\mathrm{\Gamma }}^{[n]}`$, namely $`(4.66)(4.68)`$, one finds
$`\mathrm{\Lambda }_{op}\overline{\mathrm{\Gamma }}_{n+1}^{[n]}=0,`$ (4.69)
$`{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_{n+1}^{[n]}}{\delta \lambda _a(x)}}=0,{\displaystyle \frac{\delta \overline{\mathrm{\Gamma }}_{n+1}^{[n]}}{\delta \lambda _y(x)}}=0,`$ (4.70)
$`\mathrm{\Sigma }_a(x)\overline{\mathrm{\Gamma }}_{n+1}^{[n]}=0,\mathrm{\Sigma }_y(x)\overline{\mathrm{\Gamma }}_{n+1}^{[n]}=0.`$ (4.71)
Let $`\overline{\mathrm{\Gamma }}_{n+1,div}^{[n]}`$ stand for the pole part of $`\overline{\mathrm{\Gamma }}_{n+1}^{[n]}`$. By repeating the steps going from $`(4.33)`$ to $`(4.41)`$, one can arrive at
$`\overline{\mathrm{\Gamma }}_{n+1,div}^{[n]}=\alpha _1^{(n+1)}T_{(1)}+\alpha _2^{(n+1)}T_{(2)}+\alpha _3^{(n+1)}T_{(3)}+\alpha _4^{(n+1)}T_{(4)}+\alpha _5^{(n+1)}T_{(5)},`$ (4.72)
where $`\alpha _1^{(n+1)},\mathrm{},\alpha _5^{(n+1)}`$ are constants of order $`(\mathrm{})^{n+1}`$. Therefore, in order to cancel the $`n+1`$ loop divergence the counterterm of order $`\mathrm{}^{n+1}`$ should be chosen as
$`\delta I_{count}^{[n+1]}=\overline{\mathrm{\Gamma }}_{n+1,div}^{[n]}[\psi ,\overline{\psi },W,C,\overline{C}].`$ (4.73)
Adding this counterterm, the mass term and the $`\lambda `$ terms to $`\overline{I}_{\mathrm{eff}}^{[n]}`$, one can express the effective action of order $`\mathrm{}^{n+1}`$ as
$`I_{\mathrm{eff}}^{[n+1]}[\psi ,\overline{\psi },W,C,\overline{C},\lambda ,K,L,n,l,p,n^{},l^{},p^{},g,g_1,M,m]`$
$`=I_{\mathrm{eff}}^{[0]}[\psi ^{[0]},\overline{\psi }^{[0]},W^{[0]},C^{[0]},\overline{C}^{[0]},\lambda ^{[0]},K^{[0]},L^{[0]},n^{[0]},n^{{}_{}{}^{}[0]},\mathrm{},g^{[0]},g_1^{[0]},M^{[0]},m^{[0]}],`$ (4.74)
where the bare fields and the bare parameters (to order $`(\mathrm{})^{n+1}`$) are defined as
$`W_{a\mu }^{[0]}=(Z_3^{[n+1]})^{1/2}W_{a\mu }=\left((Z_3^{[n]})^{1/2}\alpha _1^{(n+1)}\right)W_{a\mu },L_a^{[0]}=(Z_3^{[n+1]})^{1/2}L_a,`$ (4.75)
$`W_{y\mu }^{[0]}=(Z_3^{{}_{}{}^{}[n+1]})^{1/2}W_{y\mu }=\left((Z_3^{{}_{}{}^{}[n]})^{1/2}\alpha _2^{(n+1)}\right)W_{y\mu },`$ (4.76)
$`C_a^{[0]}=(\stackrel{~}{Z}_3^{[n+1]})^{1/2}C_a=\left((\stackrel{~}{Z}_3^{[n]})^{1/2}+(\alpha _3^{(n+1)}+\alpha _1^{(n+1)})\right)C_a,`$ (4.77)
$`\overline{C}_a^{[0]}=(\stackrel{~}{Z}_3^{[n+1]})^{1/2}\overline{C}_a,K_\mu ^{a[0]}=(\stackrel{~}{Z}_3^{[n+1]})^{1/2}K_\mu ^a,`$ (4.78)
$`C_y^{[0]}=(\stackrel{~}{Z}_3^{{}_{}{}^{}[n+1]})^{1/2}C_y=\left((\stackrel{~}{Z}_3^{{}_{}{}^{}[n]})^{1/2}+(\alpha _3^{(n+1)}+\alpha _2^{(n+1)})\right)C_y,`$ (4.79)
$`\overline{C}_y^{[0]}=(\stackrel{~}{Z}_3^{{}_{}{}^{}[n+1]})^{1/2}\overline{C}_y,K_\mu ^{y[0]}=(\stackrel{~}{Z}_3^{[n+1]})^{1/2}K_\mu ^y,`$ (4.80)
$`\nu _L^{[0]}=(Z_{\nu L}^{[n+1]})^{1/2}\nu _L=\left((Z_{\nu L}^{[n]})^{1/2}\alpha _4^{(n+1)}\right)\nu _L,\overline{\nu }_L^{[0]}=(Z_{\nu L}^{[n+1]})^{1/2}\overline{\nu }_L,`$ (4.81)
$`e_L^{[0]}=(Z_{eL}^{[n+1]})^{1/2}e_L=(Z_{\nu L}^{[n+1]})^{1/2}e_L,\overline{e}_L^{[0]}=(Z_{eL}^{[n+1]})^{1/2}\overline{e}_L,`$ (4.82)
$`e_R^{[0]}=(Z_{eR}^{[n+1]})^{1/2}e_R=\left((Z_{eR}^{[n]})^{1/2}\alpha _5^{(n+1)}\right)e_R,\overline{e}_R^{[0]}=(Z_{eR}^{[n+1]})^{1/2}\overline{e}_R,`$ (4.83)
$`n^{[0]}=(Z_{(n)}^{[n+1]})^{1/2}n=\left((Z_{(n)}^{[n]})^{1/2}+(\alpha _3^{(n+1)}+\alpha _4^{(n+1)})\right)n,n^{{}_{}{}^{}[0]}=(Z_{(n)}^{[n+1]})^{1/2}n^{},`$ (4.84)
$`l^{[0]}=(Z_{(l)}^{[n+1]})^{1/2}l=(Z_{(n)}^{[n+1]})^{1/2}l,l^{{}_{}{}^{}[0]}=(Z_{(l)}^{[n+1]})^{1/2}l^{},`$ (4.85)
$`p^{[0]}=(Z_{(p)}^{[n+1]})^{1/2}p=\left((Z_{(p)}^{[n]})^{1/2}\alpha _3^{(n+1)}+\alpha _5^{(n+1)}\right)p,p^{{}_{}{}^{}[0]}=(Z_{(p)}^{[n+1]})^{1/2}p^{},`$ (4.86)
$`g^{[0]}=Z_g^{[n+1]}g=(Z_3^{[n+1]})^{1/2}g,g_1^{[0]}=Z_g^{{}_{}{}^{}[n+1]}g_1=(Z_3^{{}_{}{}^{}[n+1]})^{1/2}g_1,`$ (4.87)
$`g^{[0]}=Z_g^{[n+1]}g=(Z_3^{[n+1]})^{1/2}g,g_1^{[0]}=Z_g^{{}_{}{}^{}[n+1]}g_1=(Z_3^{{}_{}{}^{}[n+1]})^{1/2}g_1,`$ (4.88)
$`M^{[0]}=Z_M^{[n+1]}M=(Z_3^{[n+1]})^{1/2}M,m^{[0]}=Z_m^{[n+1]}m=(Z_{eL}^{[n+1]})^{1/2}(Z_{eR}^{[n+1]})^{1/2}M,`$ (4.89)
and $`\lambda _a^{[0]},\lambda _y^{[0]}`$ are
$`\lambda _a^{[0]}=(Z_3^{[n+1]})^{1/2}\lambda _a,\lambda _y^{[0]}=(Z_3^{{}_{}{}^{}[n+1]})^{1/2}\lambda _y.`$ (4.90)
Therefore, in terms of such bare fields and bare parameters, $`\overline{\mathrm{\Gamma }}^{[n+1]}`$ can be expressed as
$`\overline{\mathrm{\Gamma }}^{[n+1]}[W,C,\overline{C},\psi ,\overline{\psi },K,L,n,l,p,n^{},l^{},p^{},g,g_1,M]`$
$`=\widehat{\mathrm{\Gamma }}^{[0]}[W^{[0]},C^{[0]},\overline{C}^{[0]},\psi ^{[0]},\overline{\psi }^{[0]},K^{[0]},L^{[0]},n^{[0]},n^{{}_{}{}^{}[0]},\mathrm{},g^{[0]},g_1^{[0]},M^{[0]}].`$ (4.91)
From this one can conclude that $`\overline{\mathrm{\Gamma }}^{[n+1]}`$, under $`(4.23)`$, satisfies $`(4.66)`$$`(4.68)`$ and is finite to order $`\mathrm{}^{n+1}`$. That is to say the theory is renormalisable.
V. Concluding Remarks
We have expounded that SU<sub>L</sub>(2) $`\times `$ U<sub>Y</sub>(1) electroweak theory with massive W Z fields and massive electron fields can still be quantized in a way similar to that used in Ref. by taking into account the constraint conditions caused by these mass terms and the additional condition chosen by us. We have also shown that when the $`\delta `$ functions appearing in the path integral of the Green functions and representing the constraint conditions are rewritten as Fourier integrals with Lagrange multipliers $`\lambda _a`$ and $`\lambda _y`$, the total effective action consisting of the Lagrange multipliers, ghost fields and the original fields is BRST invariant. Furthermore, with the help of the renormalisability of the theory without the the mass term of matter fields we have found the general form of the divergent part of the generating functional $`\mathrm{\Gamma }`$ and proven that the mass term of the electron fields is also harmless to the renormalisability of the theory.
It is worth while emphasizing the following special features of the SU<sub>L</sub>(2) $`\times `$ U<sub>Y</sub>(1) electroweak theory with massive W Z fields and massive electron fields. (1) These mass terms do not appear in the divergent part of $`\mathrm{\Gamma }`$. (2) The ghostโelectron coupling term $`I_m^{(C)}`$, which is caused by the mass term of the electron fields and contains the negative dimension parameter $`m/M^2`$, is not an independent term of the divergent part of $`\mathrm{\Gamma }`$. If this were not the case, the mass terms would be harmful to the renormalisability of the theory.
As pointed out in Ref. , since the whereabouts of the Higgs Bosons is still unknown, it is reasonable to ask if the successes of the standard model of the electroweak theory really depends on the Higgs mechanism and to pay attention to the theory without the Higgs mechanism.
ACKNOWLEDGMENTS
We are grateful to Professor Yang Li-ming for helpful discussions. This work was supported by National Natural Science Foundation of China and supported in part by Doctoral Programm Foundation of the Institution of Higher Education of China.
Refernces
Ze-Sen Yang, Zhining Zhou, Yushu Zhong and Xianhui Li, hep-th/9912046 7 Dec 1999.
Ze-Sen Yang, Xianhui Li, Zhining Zhou and Yushu Zhong, hep-th/9912034 5 Dec 1999.
Ze-Sen Yang, Xianhui Li, Zhining Zhou and Yushu Zhong, hep-th/0003149 17 Mar 2000.
M.Carena and C.Wagner, Phys. Rev. D37, 560(1988).
R.Delbourgo and G.Thompson, Phys. Rev. Lett. 57, 2610(1986).
M.Carena and C.Wagner, Phys. Rev. D37, 560(1988).
R.Delbourgo and G.Thompson, Phys. Rev. Lett. 57, 2610(1986).
A.Burnel, Phys. Rev. D33, 2981(1986);D33, 2985(1986);.
T.Fukuda, M.Monoa, M.Takeda and K.Yokoyama,
Prog. Theor. Phys. 66,1827(1981);67,1206(1982);70,284(1983).
S.L.Glashow, Nucl. Phys. 22, 579(1961).
L.D. Faddeev and V.N. Popov, Phys. Lett. B25, 29(1967).
B.S. De Witt, Phys. Rev. 162,1195,1239(1967).
L.D. Faddev and A.A.Slavnov, Gauge Field: Introduction to Quatum Theory,
The Benjamin Cummings Publishing Company, 1980.
G.H.Lee and J.H.Yee, Phys. Rev. D46, 865(1992).
C.Itzykson and F-B.Zuber, Quantum Field Theory, McGraw-Hill, New York, 1980.
Yang Ze-sen, Advanced Quantum Machanics, Peking University Press, 2-ed. Beijing, 1995. |
warning/0003/hep-ph0003312.html | ar5iv | text | # Baryogenesis vs. proton stability in theories with extra dimensions
## I Introduction
Despite the great success of Quantum Field Theory, a consistent scenario where gravity is also included still lacks. The most promising framework that could help in this task is string theory, whose consistency requires additional dimensions beyond the standard $`3+1`$. This extra space is usually assumed to be compact, with a small compactification radius of order $`M_p^{\mathrm{\hspace{0.17em}1}}`$. However, it has been observed in ref. that, having no test of gravity below the millimeter scale, we do not really need such a tiny compactification radius, provided the extra dimensions are accessible only to gravitational interactions. The Standard Model degrees of freedom must indeed be localized on a $`3`$ dimensional wall whose inverse thickness does not exceed the scale of energy, of order TeV , we currently probe in accelerator experiments.
The choice of such large compactification scale has the main goal of solving (or at least of weakening) the hierarchy problem. Denoting by $`V_n`$ the volume of the compact space โ assumed in to have a trivial metric โ and by $`M`$ the fundamental scale of gravity, the observed Planck mass is obtained by the relation $`M_p^2=M^{n+2}V_n`$. Under the condition $`n2`$, $`M`$ can be safely assumed to be very close to the electroweak scale, without conflicting with either cosmological, astrophysical, or laboratory bounds.
Although considerably improving the standard situation, the above scenario retains however some degree of fine tuning, connected to the largeness of the quantity $`V_nM^n`$. A better result in this regard is provided by the more recent work , where, due to the presence of cosmological constants in the bulk and on two walls, the metric is nonfactorizable with an exponential scaling in the extra space. This fact allows the achievement of both a (phenomenologically) acceptable theory with just one extra dimension (that could be even infinite in extension ) and a more satisfactory solution to the hierarchy problem.
There are some other aspects common to all of these theories besides the ones discussed above. In particular, both proton stability and baryogenesis may be problematic in models with very low fundamental masses.
For what concerns proton stability in Grand Unified Theories, the standard way to achieve it is to increase the mass of the additional bosons up to about $`10^{15}10^{16}`$GeV . In the framework of theories with extraโdimensions, an interesting mechanism has been suggested in ref. . <sup>*</sup><sup>*</sup>*See also for alternative suggestions. In this paper, a dynamical mechanism for the localization of fermions on the wall is adopted: leptons and quarks are however localized at two slightly displaced positions in the extra space, and this naturally suppresses the interactions which โconvertโ the latter in the former.
However, the observed baryon asymmetry requires baryon number ($`B`$) violating interactions to have been effective in the first stages of the evolution of the Universe. In this paper we thus wonder how this last requirement can be satisfied in a theory which adopts the idea of , to ensure proton stability now and baryon production in the past. Our proposal is that thermal corrections, which are naturally relevant at early times, may modify the localization of quarks and leptons so to weaken the mechanism that suppresses the $`B`$ violating interactions. There exist other proposals for baryogenesis in these theories : in the work , after considering several bounds on baryogenesis with large extra dimensions, a mechanism based on nonrenormalizable operators is proposed; in ref. baryon number is violated by โevaporationโ of brane bubbles that carry a net baryonic charge into the bulk, and the matterโantimatter asymmetry can be due to a primordial collision of our brane with another one, that carried away the missing antimatter; in ref. baryogenesis is obtained via leptogenesis, the latter being due to the existence of sterile neutrinos in the bulk.
The plan of the work is the following. In section $`2`$ we review the mechanism used in ref. to localize chiral fermions on a domain wall and to suppress the rate of baryon number violating interactions. In our work, however, we also take into account the finite thickness of the wall: this enforces on the parameters of the model some bounds which are stronger than the ones reported in ref. . In section $`3`$ we estimate the thermal corrections to the parameters of the theory, and in particular to the function that measures the suppression of the B violating interactions. Being our model nonrenormalizable, a perturbative treatment can be meaningful only at low energies. Anyhow, it is conceivable that the results we present in section $`3`$ can be a hint for the behavior of the theory at higher temperatures. The issue of baryogenesis is faced in section $`4`$, where we consider a very simple example reminiscent of GUT baryogenesis. In this mechanism, the baryon asymmetry is achieved through the dacay of a boson, whose interactions violate baryon number. In order for the model to work, the boson must be out of equilibrium before decaying, and this is not obvious in theories with low fundamental masses. In those theories, the Hubble parameter is indeed very low at energies below the fundamental scale of gravity, that sets the natural cutoff of the theory. We will thus consider a model where the bosons responsible for the baryon asymmetry are produced nonthermally (for instance, at the end of inflation) fulfilling thus naturally the out of equilibrium requirement. After considering other bounds, such as the stability of the kink under thermal corrections, we finally calculate the baryon asymmetry in a $`BL`$ conserving scheme. In the conclusions we discuss our results and their possible future extensions.
## II Localization
A simple mechanism for localizing fermions on a wall has been recently revisited in ref. .
In this paper the idea is illustrated in the easiest case where only one extra dimension is added to the usual four. The main ingredient that is needed is a scalar field $`\varphi `$ which couples to the fermionic field $`\psi `$ through the full five dimensional Yukawa interaction $`g\varphi \overline{\psi }\psi `$ and whose expectation value $`\varphi `$ varies along the extra dimension, but it is constant on our four-dimensional world. In this way the VEV $`\varphi `$ breaks the full translational invariance, as it is needed to have a preferred direction orthogonal to the wall.
It is possible to show that in this case the fermionic field localizes where its total mass $`m=m_0+g\varphi `$ ($`m_0`$ is the bare fermionic mass in the five dimensional theory) vanishes, i.e. on a wall with three spatial dimensions characterized by a particular position $`x_5`$ in the transverse direction.
For definiteness, we consider the theory described by the lagrangian
$`_{\varphi \psi }`$ $`=`$ $`\overline{\psi }(i/_5+{\displaystyle \frac{1}{\stackrel{~}{M}_0^{1/2}}}\varphi \left(y\right)+m_0)\psi `$ (1)
$`_\varphi `$ $`=`$ $`{\displaystyle \frac{1}{2}}_\mu \varphi ^\mu \varphi \left(\mu _0^2\varphi ^2+\lambda _0\varphi ^4\right),`$ (2)
where $`yx_5`$ is the fifth coordinate, the fields and the parameters have the following mass dimensions
$$\left[\varphi \right]=3/2,\left[\psi \right]=2,\left[m_0\right]=\left[\mu _0\right]=\left[\stackrel{~}{M}_0\right]=1,\left[\lambda _0\right]=\mathrm{\hspace{0.17em}1},$$
(3)
and where the suffix $`0`$ indicates the value of the parameters at zero temperature.
As we said, the localization position of the fermions depends on the vacuum configuration of the field $`\varphi `$. If we consider the kink solution
$$\varphi =\frac{\mu _0}{\sqrt{2\lambda _0}}\mathrm{tanh}\left(\mu _0y\right),$$
(4)
and we approximate it with a straight line interpolating between the two vacua (see figure $`1`$)
$`\{\begin{array}{cc}\varphi \left(y\right)\frac{\mu _0^2}{\sqrt{2\lambda _0}}y,|y|<\frac{1}{\mu _0}\hfill & \\ \varphi \left(y\right)\pm \frac{\mu _0^2}{\sqrt{2\lambda _0}},|y|>\frac{1}{\mu _0},\hfill & \end{array}`$ (5)
we see that the localization can occur only if
$$m_0<\frac{\mu _0}{\sqrt{2\lambda _0\stackrel{~}{M}_0}},$$
(6)
since otherwise the total fermion mass
$$m_{\mathrm{tot}}=\frac{1}{\stackrel{~}{M}_0^{1/2}}\varphi \left(y\right)+m_0$$
(7)
never vanishes.
It can be shown that, from the four dimensional point of view, a left handed chiral massless fermionic field results from the localization mechanism, if the above configuration (4) is assumed for the scalar $`\varphi `$. The right handed part remains instead delocalized in the whole space. This is not a problem since it is customary to limit the Standard Model and the MSSM fermionic content only to left handed fields. <sup>ยง</sup><sup>ยง</sup>ยงConcerning the cancellation of anomalies on the wall and recovering the Standard Model running of the coupling constants, see . The right handed fields can also be localized if a kinkโantikink solution is assumed for the scalar $`\varphi `$. As a result, the left fields continue to be localized on the kink, while the right ones are confined to the antikink. If the kink and the antikink are sufficiently far apart, the left handed and right handed fermions however do not interact and again the model reproducing our four dimensional world must be built by fermions of a defined chirality. The fermionic content of the full dimensional theory is in this case doubled with respect to the usual one, and observers on one of the two walls will refer to the other as to a โmirror worldโ. The presence of this kinkโantikink configuration may be required by stability consideration if thermal effects are considered, as it is the case in the next sections. However most of the physics in one brane is not affected by the presence of the mirror one, and in the most of the present work we will concentrate on a single wall as if only the kink (4) configuration was present.
In order to give mass to the fermions, some other scalar field acting as a Higgs in the four dimensional theory must be considered. As it is shown in ref. , the mechanism described above could give an explanation to the hierarchy among the Yukawa couplings responsible for the fermionic mass matrix. If indeed one chooses different five dimensional bare masses for the different fermionic fields, the latter are localized at different positions in the fifth direction. As a consequence, the wave functions of different fermions do only partially overlap, and increasing the difference between the five dimensional bare masses of two fermions results in suppressing their mutual interactions.
The same idea can be adopted to guarantee proton stability. Let us give, respectively, leptons and baryons the โmassesโ
$$\left(m_0\right)_l=0,\left(m_0\right)_b=m_0,$$
which correspond to the localizations The last inequality in the next expression comes from (6). We assume quarks of different generations to be located in the same $`y`$ position in order to avoid dangerous FCNC mediated by the Kaluza-Klein modes of the gluons .
$$y_l=0,y_b=\frac{m_0\sqrt{2\lambda _0\stackrel{~}{M}_0}}{\mu _0^2}<\frac{1}{\mu _0}$$
The shape of the fermion wave functions along the fifth dimension can be cast in an explicit and simple form if we consider the limit $`y_b1/\mu _0`$, in which the effect of the plateau for $`y>1/\mu _0`$ can be neglected: This is also the limit in which the approximation (5) is valid.
$`f_l\left(y\right)`$ $`=`$ $`\left({\displaystyle \frac{\mu _0^2}{\sqrt{2\lambda _0\stackrel{~}{M}_0}\pi }}\right)^{1/4}\mathrm{exp}\left\{{\displaystyle \frac{\mu _0^2y^2}{2\sqrt{2\lambda _0\stackrel{~}{M}_0}}}\right\}`$ (8)
$`f_b\left(y\right)`$ $`=`$ $`\left({\displaystyle \frac{\mu _0^2}{\sqrt{2\lambda _0\stackrel{~}{M}_0}\pi }}\right)^{1/4}\mathrm{exp}\left\{{\displaystyle \frac{\mu _0^2\left(yy_b\right)^2}{2\sqrt{2\lambda _0\stackrel{~}{M}_0}}}\right\}.`$ (9)
We assume the Standard Model to be embedded in some theory which, in general, contains some additional bosons $`X`$ whose interactions violate baryon number conservation. If it is the case, the four fermion interaction $`qqql`$ can be effectively described by
$$d^4x๐y\frac{qqql}{\mathrm{\Lambda }m_X^2},$$
(10)
where $`m_X`$ is the mass of the intermediate boson $`X`$ and $`\mathrm{\Lambda }`$ is a parameter of mass dimension one related to the five-dimensional coupling of the X-particle to quarks and leptons.
This scattering is thus suppressed by
$`I`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Lambda }m_X^2}}{\displaystyle ๐y\frac{\mu _0^2}{\pi \sqrt{2\lambda _0\stackrel{~}{M}_0}}\mathrm{exp}\left\{\frac{\mu _0^2/2}{\sqrt{2\lambda _0\stackrel{~}{M}_0}}\left[y^2+3\left(yy_b\right)^2\right]\right\}}=`$ (11)
$`\mathrm{}`$ $`=`$ $`{\displaystyle \frac{\mu _0}{\mathrm{\Lambda }m_X^2\sqrt{2\pi }\left(2\lambda _0\stackrel{~}{M}_0\right)^{1/4}}}\mathrm{exp}\left\{{\displaystyle \frac{3\left(2\lambda _0\stackrel{~}{M}_0\right)^{1/2}}{8}}{\displaystyle \frac{m_0^2}{\mu _0^2}}\right\}.`$ (12)
Current proton stability requires I
<
(1016GeV)โ2,
<
๐ผsuperscriptsuperscript1016GeV2I\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\left(10^{16}\,\mbox{GeV}\right)^{-\,2}\;, that is
m0ฮผ0
>
2006Log10(ฮmX2ฮผ0/GeV2)(2ฮป0M~0)1/4.
>
subscript๐0subscript๐02006subscriptLog10ฮsuperscriptsubscript๐๐2subscript๐0superscriptGeV2superscript2subscript๐0subscript~๐014\frac{m_{0}}{\mu_{0}}\mathrel{\vbox{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}}\frac{\sqrt{200-6\,\mbox{Log}_{10}\left(\frac{\Lambda\,m_{X}^{2}}{\mu_{0}}\Big{/}\mbox{GeV}^{2}\right)}}{\left(2\,\lambda_{0}\,{\widetilde{M}}_{0}\right)^{1/4}}\;\;. (13)
The numerator in the last equation is quite insensitive to the mass scales of the model, and โ due to the logarithmic mild dependence โ can be safely assumed to be of order $`10`$. For definiteness, we will thus fix it at the value of $`10`$ in the rest of our work.
Conditions (6) and (13) give altogether
10ฮผ0(2ฮป0M~0)1/4
<
m0
<
ฮผ0(2ฮป0M~0)1/2,
<
10subscript๐0superscript2subscript๐0subscript~๐014subscript๐0
<
subscript๐0superscript2subscript๐0subscript~๐012\frac{10\,\mu_{0}}{\left(2\,\lambda_{0}\,{\widetilde{M}}_{0}\right)^{1/4}}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}m_{0}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\frac{\mu_{0}}{\left(2\,\lambda_{0}\,{\widetilde{M}}_{0}\right)^{1/2}}\;\;, (14)
that we can rewrite
{2ฮป0M~0
<
104m0ฮผ0
>
102.cases
<
2subscript๐0subscript~๐0superscript104otherwise
>
subscript๐0subscript๐0superscript102otherwise\displaystyle\cases{2\,\lambda_{0}\,{\widetilde{M}}_{0}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10^{-4}&\cr\frac{m_{0}}{\mu_{0}}\mathrel{\vbox{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}}10^{2}\;\;.} (15)
The last limit in eqs. (15) is stronger than the one given in ref. where proton stability is satisfied if the ratio of the massive scales of the model is of order $`10`$. However, in ref. the field $`\varphi `$ simply scales linearly as a function of $`y`$, while we expect that whenever a specific model is assumed, conditions analogous to our (6) and (15) should be imposed.
## III Thermal correction to the coefficients
Once the localization mechanism is incorporated in a low energy effective theory โ as the system (1) may be considered โ, one can legitimately ask if thermal effects could play any significant role. In the present work we are mainly interested in any possible change in the argument of the exponential in eq. (11), that will be the most relevant for the purpose of baryogenesis. For this reason, we introduce the dimensionless quantity
$$a(T)=\frac{m(T)^2}{\mu (T)^2}\sqrt{2\lambda (T)\stackrel{~}{M}(T)}.$$
(16)
From eqs.(13) and (15), we can set a(0)
>
100
>
๐0100a(0)\mathrel{\vbox{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}}100 at zero temperature. Thermal effects will modify this value. There are however some obstacles that one meets in evaluating the finite temperature result. Apart from some technical difficulties arising from the fact that the scalar background is not constant, the main problem is that nonperturbative effects may play a very relevant role at high temperature. As it is customary in theories with extra dimensions, the model (1) is nonrenormalizable and one expects that there is a cut-off (generally related to the fundamental scale of gravity) above which it stops holding. Our considerations will thus be valid only for low temperature effects, and may only be assumed as a rough indication for what can happen at higher temperature.
Being aware of these problems, by looking at the dominant finite-temperature one-loop effects, we estimate the first corrections to the relevant parameters to be
$`\{\begin{array}{cc}\lambda \left(T\right)=\lambda _0+c_\lambda \frac{T}{\stackrel{~}{M}_0^2}\hfill & \\ \stackrel{~}{M}\left(T\right)=\stackrel{~}{M}_0+c_{\stackrel{~}{M}}T\hfill & \\ m\left(T\right)=m_0+c_m\frac{T^2}{\stackrel{~}{M}_0}\hfill & \\ \mu ^2\left(T\right)=\mu _0^2+c_\mu \frac{T^3}{\stackrel{~}{M}_0},\hfill & \end{array}`$ (17)
where the $`c`$โs are dimensionless coefficients whose values are related to the exact particle content of the theory.
In writing the above equations, the first of conditions (15) has also been taken into account. For example, both a scalar and a fermionic loop contribute to the thermal correction to the parameter $`\lambda _0`$. While the contribution from the former is of order $`\lambda _0^2T`$, the one of the latter is of order $`T/\stackrel{~}{M}_{0}^{}{}_{}{}^{2}`$ and thus dominates. <sup>\**</sup><sup>\**</sup>\**Notice also that in our model loops with internal leptons dominate over loops with internal quarks, since the former have vanishing five dimensional bare mass and thus are not Boltzmann suppressed. However, although this choice is the simplest one, one may equally consider the most general case where all the fermions have a nonvanishing five dimensional mass.
Substituting eqs. (17) into eq. (16), we get, in the limit of low temperature,
$$a\left(T\right)a\left(0\right)\left[1+\frac{T}{\stackrel{~}{M}_0}\left(\frac{c_\lambda }{2\lambda _0\stackrel{~}{M}_0}+\frac{c_{\stackrel{~}{M}}}{2}+\frac{2c_mT}{m_0}\frac{c_\mu T^2}{\mu _0^2}\right)\right].$$
(18)
From the smallness of the quantity $`\lambda _0\stackrel{~}{M}_0`$ \[see cond. (15)\] we can safely assume (apart from high hierarchy between the $`c`$โs coefficients that we do not expect to hold) that the dominant contribution in the above expression comes from the term proportional to $`c_\lambda `$.
We thus simply have
$$a\left(T\right)a\left(0\right)\left(1+c_\lambda \frac{T}{2\lambda _0\stackrel{~}{M}_0^2}\right).$$
(19)
We notice that the parameter $`c_\lambda `$, being related to the thermal corrections to the $`\varphi ^4`$ coefficient due to a fermion loop, is expected to be negative : the first thermal effect is to decrease the value of the parameter $`a(T)`$, making hence the baryon number violating reactions more efficient at finite rather than at zero temperature.
There is another effect which may be very crucial at finite temperature, linked to the stability of the $`Z_2`$ symmetry. When a temperature is turned on, we generally expect the formation of a fermionโantifermion condensate $`\overline{\psi }\psi 0`$. If it is the case, the Yukawa coupling $`\varphi \overline{\psi }\psi `$ in the lagrangian (1) renders one of the two vacua unstable. While this leads to an instantaneous decay of the kink configuration, a kinkโantikink system could have a sufficiently long lifetime provided the two objects are enough far apart.
## IV Baryogenesis
We saw in the previous section that thermal effects may increase the rate of baryon number violating interactions of our system. This is very welcome, since a theory which never violates baryon number cannot lead to baryogenesis and thus cannot reproduce the observed Universe. Anyhow baryon number violation is only one of the ingredients for baryogenesis, and the aim of this section is to investigate how the above mechanism can be embedded in a more general context.
A particular scheme which may be adopted is baryogenesis through the decay of massive bosons $`X`$. <sup>โ โ </sup><sup>โ โ </sup>โ โ We may think of these bosons as the intermediate particles which mediate the four fermion interaction described by the term (10). This scheme closely resembles GUT baryogenesis, but there are some important peculiarities due to the different scales of energy involved. In GUT baryogenesis the massive boson $`X`$, coupled to matter by the interaction $`gX\psi \overline{\psi }`$, has the decay rate
$$\mathrm{\Gamma }\alpha m_x,\alpha =\frac{g^2}{4\pi }.$$
(20)
An important condition is that the $`X`$ boson decays when the temperature of the Universe is below its mass (out of equilibrium decay), in order to avoid thermal regeneration. From the standard equation for the expansion of the Universe,
$$Hg_{}^{1/2}\frac{T^2}{M_p}$$
(21)
(where $`g_{}`$ is the number of relativistic degrees of freedom at the temperature $`T`$), this condition rewrites
mX
>
g1/2ฮฑMp.
>
subscript๐๐superscriptsubscript๐12๐ผsubscript๐๐m_{X}\mathrel{\vbox{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}}g_{*}^{-1/2}\,\alpha\,M_{p}\;\;. (22)
If $`X`$ is a Higgs particle, $`\alpha `$ can be as low as $`10^{\mathrm{\hspace{0.17em}6}}`$. Even in this case however the $`X`$ boson must be very massive. In principle this may be problematic in the theories with extra dimensions we are interested in, which have the main goal of having a very low fundamental scale.
There are some possibilities to overcome this problem. One is related to a possible deviation of the expansion of the Universe from the standard behavior. The issue of the Friedmann law in models with large extra dimensions has been indeed subject of intense debate in the recent past. In the work a detailed analysis of the Einstein equations with one extra dimension, shows that the expansion rate $`H`$ should (in absence of any energy in the bulk) be proportional to the energy density $`\rho `$ on the brane. This behavior strongly conflicts with the standard one $`H\rho ^{1/2}`$. In refs. it was then shown that the standard expansion law could be achieved, at least at low temperatures, by a suitable fine tuning of the vacuum energies in the Universe. In particular, this is the case for the RandallโSundrum model , which offers one of the most satisfactory solutions to the hierarchy problem. The solution proposed in these works is however itself plagued by some other cosmological problems. For example, in the RS model gravity turns out to be repulsive on our brane. All the above problems are overcome when some mechanism for the stabilization of the radion is taken into account, as the analyses show. In particular, in ref. , the case of the model is examined, and both the standard Friedmann law and the โcorrectโ sign for the Newton constant are obtained. The analysis of ref. is however performed by computing only the first-order term of the expansion of the square of the Hubble parameter $`H^2`$ as a power series of the energy density $`\rho `$. Terms of order $`\rho ^2`$ could become relevant at temperatures above $`1\text{TeV}`$ or so . This may result in an accelerated expansion of the Universe at high temperatures, and the out of equilibrium condition for the $`X`$ bosons could be consequently considerabily favoured.
However, both the facts that we do not know the exact behavior of the expansion rate at high temperatures in the model , and that one may be interested in embedding our baryogenesis mechanism in some other cosmological scenario, lead us to discuss alternative solutions for the out of equilibrium problem. One very natural possibility is to create the $`X`$ particles non thermally and to require the temperature of the Universe to be always smaller than their mass $`m_X`$. In this way, one kinematically forbids regeneration of the $`X`$ particles after their decay. In addition, although interactions among these bosons can bring them to thermal equilibrium, chemical equilibrium cannot be achieved.
Nonthermal creation of matter has raised a considerable interest in the last years. In particular, it has been shown that this production can be very efficient during the period of coherent oscillations of the inflaton field after inflation . The efficiency of this mechanism has also been exploited in the work to revive GUT baryogenesis in the context of standard four dimensional theories. Here, we will not go into the details of the processes that could have lead to the production of the $`X`$ bosons. Rather, we will simply assume that, after inflation, their number density is $`n_X`$. To simplify our computations, we will also suppose that their energy density dominates over the thermal bath produced by the perturbative decay of the inflaton field.
Just for definiteness, let us consider a very simple model where there are two species of $`X`$ boson which can decay into quarks and leptons, according to the four dimensional effective interactions
$$gX\overline{q}\overline{q},ge^{a/4}Xlq,$$
(23)
where (remember the suppression given by the different localization of quarks and leptons) the quantity $`a`$ is defined in eq. (16). Again for definiteness we will consider the minimal model where no extra fermionic degrees of freedom are added to the ones present in the Standard Model. Moreover we will assume $`BL`$ to be conserved, even though the extension to a more general scheme can be easily performed.
The decay of the $`X`$ bosons will reheat the Universe to a temperature that can be evaluated to be
$$T_{\mathrm{rh}}\left(\frac{30}{\pi ^2}\frac{m_Xn_X}{g_{}}\right)^{1/4}.$$
(24)
Since we do not want the $`X`$ particles to be thermally regenerated after their decay, we require Trh
<
mX
<
subscript๐rhsubscript๐๐T_{\mathrm{rh}}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}m_{X}, that can be rewritten as an upper bound on $`n_X`$
nX
<
30(g100)mX3.
<
subscript๐๐30subscript๐100superscriptsubscript๐๐3n_{X}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}30\,\left(\frac{g_{*}}{100}\right)\,m_{X}^{3}\,\,. (25)
Another limit comes from the necessity to forbid the $`B`$ violating four fermion interaction (10) to erase the $`B`$ asymmetry that has been just created by the decay of the $`X`$ bosons. We thus require the interaction (10) to be out equilibrium at temperatures lower than $`T_{\mathrm{rh}}`$. From eq. (11) we see that we can parametrize the four fermion interaction with a coupling $`g^2e^{3a/8}/m_X^2`$. Hence, the out of equilibrium condition reads
g4e3a/4
<
gmXMp(mXTrh)3.
<
superscript๐4superscript๐3๐4subscript๐subscript๐๐subscript๐๐superscriptsubscript๐๐subscript๐rh3g^{4}\,e^{-3\,a/4}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}g_{*}\,\frac{m_{X}}{M_{p}}\,\left(\frac{m_{X}}{T_{\mathrm{rh}}}\right)^{3}\;\;. (26)
One more upper bound on the reheating temperature comes from the out of equilibrium condition for the sphalerons. This requirement is necessary only if one chooses the theory to be $`BL`$ invariant, while it does not hold for $`BL`$ violating schemes. We can approximately consider the sphalerons to be in thermal equilibrium at temperatures above the electroweak scale. Thus, if $`BL`$ is a conserved quantity, we will require the reheat temperature to be smaller than about $`100\mathrm{GeV}`$.
If one neglects the presence of the thermal bath prior to the decay of the $`X`$ bosons, the very first decays will be only into couples of quarks, since the channel into one quark and one lepton is strongly suppressed by the $`e^{a\left(T=0\right)}`$ factor due to the fact that the kink is not modified by any thermal correction. However, the decay process is not an instantaneous event. It is shown in ref. that the particles produced in the very first decays are generally expected to thermalize very rapidly, so to create a thermal bath even when most of the energy density is still stored in the decaying particles. <sup>โกโก</sup><sup>โกโก</sup>โกโกAs shown in ref. , what is called the reheating temperature is indeed the temperature of the thermal bath when it starts to dominate. After the first decays, the temperature of the light degrees of freedom can be even much higher than $`T_{\mathrm{rh}}`$. The temperature of this bath can even be considerably higher than the final reheating themperature. The presence of the heat bath modifies in turn the shape of the kink, as shown in the previuos section, and we can naturally expect that this modification enhances the $`B`$ violating interaction.
If the energy density of the Universe is dominated by the $`X`$ bosons before they decay, one has
$$\eta _B\mathrm{\hspace{0.17em}0.1}\left(N_XT_{\mathrm{rh}}/m_X\right)r\overline{r},$$
(27)
where $`N_X`$ is the number of degrees of freedom associated to the $`X`$ particles and $`r\overline{r}`$ is the difference between the rates of the decays $`Xql`$ and $`\overline{X}\overline{q}\overline{l}`$.
We denote with $`X_1`$ and $`X_2`$ the two species of bosons whose interactions (23) lead to baryon number violation, and parametrize by $`ฯต`$ the strength of CP-violation in these interactions. Considering that $`e^{2a}`$ is always much smaller than one, we get
$$r\overline{r}3g^2e^{a/2}ฯต\mathrm{I}m\text{I}_{SS}\left(M_{X_1}/M_{X_2}\right),$$
(28)
where the function $`\mathrm{I}m\text{I}_{SS}(\rho )=[\rho ^2\mathrm{L}og(1+1/\rho ^2)1]/(16\pi )`$ can be estimated to be of order $`10^310^2`$. It is also reasonable to assume $`ฯต10^21`$.
Collecting all the above estimates, and assuming $`N_X`$ to be of order $`10`$, we get
$$\eta _B\left(10^5\mathrm{\hspace{0.17em}10}^2\right)g^2\frac{T_{\mathrm{rh}}}{m_X}e^{a\left(T_{\mathrm{rh}}\right)/2}.$$
(29)
From the requirement Trh
<
mX
<
subscript๐rhsubscript๐๐T_{\mathrm{rh}}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}m_{X} we get an upper limit on the baryon asymmetry
ฮทB
<
(105โ102)g2ea/2,
<
subscript๐๐ตsuperscript105superscript102superscript๐2superscript๐๐2\eta_{B}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\left(10^{-5}\,-\,10^{-2}\right)\,g^{2}\,e^{-a/2}\,\,, (30)
where the factor $`a\left(T\right)`$ has to be calculated for a value of $`T`$ of the order of the reheating temperature.
We get a different limit on $`\eta _B`$ from the bound (26): assuming $`m_X\text{TeV}`$ and $`g_{}100`$ indeed one obtains
ฮทB
<
(106โ1010)g2/3ea/4.
<
subscript๐๐ตsuperscript106superscript1010superscript๐23superscript๐๐4\eta_{B}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\left(10^{-6}\,-\,10^{-10}\right)\,g^{2/3}\,e^{-a/4}\,\,. (31)
Since the observed amount of baryon asymmetry is of order $`10^{10}`$, even in the case of maximum efficiency of the process (that is, assuming maximal $`CP`$ violation and $`g1`$), we have that both bounds (30) and (31) imply that $`a\left(T_{\mathrm{rh}}\right)`$ has to be smaller than about $`40`$.
Unfortunately, the temperature at which the condition a(T)
<
40
<
๐๐40a\left(T\right)\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}40\> occurs cannot be evaluated by means of the expansion of eq. (19), that have been obtained under the assumption $`\left|a\left(T\right)a\left(0\right)\right|a\left(0\right)`$. On the other hand, it is remarkable that our mechanism may work with a ratio $`a\left(T_d\right)/a\left(0\right)`$ of order one. We thus expect that a successful baryogenesis may be realized for a range of the parameters of our theory which โ although not evaluable through a perturbative analysis โ should be quite wide and reasonable.
In scenarios with large extra dimensions and low scale gravity, the maximal temperature reached by the Universe after inflation is strongly bounded from above in order to avoid overproducing Kaluza-Klein graviton modes, which may eventually contradict cosmological observations . For instance, in models with two large extra dimensions the reheating temperature has to be less than about $`10\text{MeV}`$. This value would be too low for our scenario since $`\eta _B`$ is proportional to the ratio $`T_{\mathrm{rh}}/m_X`$, and hence the observed amount of baryons would be reproduced at the price of an unnaturally small value of $`a\left(T_{\mathrm{rh}}\right)`$. However, other schemes with extra dimensions exist where the bounds on $`T_{\mathrm{rh}}`$ are less severe. For example, in the proposals the mass of the first graviton KK mode is expected to be of order TeV. The reheating temperature can thus safely be taken to be of order $`10100\text{GeV}`$.
An alternative way to overcome the bound (22) relies on the fact that, as observed in the work , the maximal temperature reached by the thermal bath during reheating can indeed be much higher than the final reheating temperature. In this case, even if $`T_{\mathrm{rh}}`$ is considerably lower than $`m_X`$, $`X`$ particles can be produced in a significant amount, and the out of equilibrium condition is easily achieved. However, the treatment of this mechanism is in our case somewhat different from the one given in ref. : due to the slowness of the expansion of the Universe, the $`X`$ bosons will decay before the freeze out of their production. The final baryon asymmetry cannot be estimated with the use of the formulae of , which are valid only if the decay of the $`X`$ particles occurs well after their freeze out.
There are of course several possible baryogenesis schemes alternative to the one just presented. A possible option which also requires a minimal extension to the Standard Model could be to achieve the baryon asymmetry directly through the $`4`$ fermions interactions $`q+qq+l`$ in the thermal primordial bath. The out of equilibrium condition may be provided by the change of the kink as the temperature of the bath decreases. <sup>\**</sup><sup>\**</sup>\**This condition may be easily achieved due of the exponential dependence of the rate of this process on the temperature, see eq. (11). What may be problematic is the source of $`CP`$ violation which may lead the creation of the baryon asymmetry. A possibility in this regard may be provided by considering a second Higgs doublet, but the whole mechanism certainly deserves a deep analysis on its own.
## V Conclusions
The present work concerns the important issue of baryogenesis in theories with large extraโdimensions. Since the observed proton stability requires to a very high degree of accuracy baryon conservation at zero temperature, this task may be problematic within the above theories, which have very low fundamental scales.
Our proposal relies on the localization mechanism for fermions discussed in ref. . While in this work the present proton stability is due to a different localization (in the transverse direction) of leptons and quarks, we believe that thermal corrections may activate early baryon violating interactions.
In our work we first provide a general discussion of the above scheme, without referring to any particular mechanism of baryogenesis. We find indeed that the first thermal corrections are in the direction of increasing the rate of baryon violations.
We then consider a very specific example, where the matterโantimatter asymmetry is achieved through the decay of a (relatively) heavy boson in a $`BL`$ conserving context. In this situation the Sakharov out of equilibrium condition can be obtained in the simplest way by considering nonthermal production of the bosons responsible for $`B+L`$ violation.
Several bounds apply to the whole mechanism. The most general ones concern the localization procedure (we have found that the limits given in ref. become more stringent once the thickness of the wall is considered) and its stability against thermal corrections. In addition, there are some other constraints which hold in the particular scheme of baryogenesis we adopted. The temperature of the heat bath right after the production of the baryon asymmetry cannot be too high, to avoid thermal regeneration of the bosons that induced baryogenesis. Moreover, this temperature has not to exceed the electroweak scale, in order not to activate the sphaleron transitions that would erase the $`B+L`$ asymmetry produced at some higher energy. Of course, this last bound can be easily overcome by considering some $`BL`$ nonconserving process.
We have found that the observed baryonic asymmetry can be accomplished quite naturally in our example, and we believe this should be the case in a more general context as well.
Possible extensions of the present work are related to the generality of the scenario we discussed. Our idea indeed relies only on the localization mechanism adopted in and not on the geometry of the bulk, nor on the details of the interactions responsible for the baryon asymmetry.
Future works could thus proceed in two directions. Firstly, one could try to embed the scheme here described in a more complete cosmological setting. Secondly, some other baryogenesis mechanisms, for example ร la AffleckโDine (which does not require very high energy scales), may be explored.
###### Acknowledgements.
We thank J. MarchโRussell, G. Mussardo, H.P. Nilles, M. Pietroni, and A. Riotto for interesting and stimulating discussions. This work is partially supported by the MURST research project โAstroparticle Physicsโ. |
warning/0003/hep-th0003091.html | ar5iv | text | # A NEW BASIS FUNCTION APPROACH TO โT HOOFT EQUATION
## 1 Introduction
It is expected that light-front (LF) quantization provides an powerful tool for studying many-body relativistic field theories . The bare vacuum is equal to the physical vacuum in the LF coordinate, since all constituents must have non-negative longitudinal momentum. This simple structure of the true vacuum enables us to avoid the insuperable problems which appeared in the Tamm-Dancoff (TD) approximation in the equal time frame. Therefore, the TD approximation is commonly used in the context of the LF quantization .
Many authors <sup>-</sup> have developed the effectual techniques for solving LFTD equations in several models. Bergknoff first applied LFTD approximation to the massive Schwinger model model , which is the extension of the simplest (1+1)-dimensional QED<sub>2</sub> . He obtained the so-called Bergknoff equation, which is the light front Einstein-Schrรถdinger equation truncated to one fermion-antifermion pair. Mo and Perry presented a prevailing method to handle the ground state and the excited state in the massive Schwinger model. They concluded that the Jacobi polynomials are appropriate basis functions to study the massive Schwinger model. Harada and his coworker investigated the massive Schwinger model with $`\mathrm{SU}(2)`$ flavor symmetry, including up to four fermion sectors. They applied of simpler basis functions, which are essentially equivalent to the Jacobi polynomials, to the model. Sugihara and collaborators studied 2-dimensional $`\mathrm{SU}(\mathrm{N}_\mathrm{C})`$ Quantum ChromoDynamics(QCD) , including four fermion sectors, by means of the basis functions of Harada et al.
Although excellent papers exist concerning massless and massive Schwinger models and 2-dimensional QCD <sup>-</sup>, including excited states, it is worth analyzing the โโt Hooftโ equation, the lowest order LFTD equation in $`\mathrm{SU}(\mathrm{N}_\mathrm{C})`$ gauge theory. This is because there is a mathematical interest in the basis function method. There is no mathematical evidence that the wave function can be expanded in terms of the conventional basis function. Contrarily, there is an evidence that the conventional method breaks down if we try to improve the approximation. We would therefore like to improve the basis functions so as to avoid such difficulties.
## 2 Conventional Basis function Method
The โt Hooft equation for two dimensional $`\mathrm{SU}(\mathrm{N}_\mathrm{C})`$ gauge theory is given in the form
$`M^2\mathrm{\Phi }(x)`$ $`=`$ $`{\displaystyle _{1/2}^{1/2}}๐yH(x,y)\mathrm{\Phi }(y)`$ (1)
$``$ $`{\displaystyle \frac{4(m^21)}{14x^2}}\mathrm{\Phi }(x)\mathrm{}{\displaystyle _{1/2}^{1/2}}๐y{\displaystyle \frac{\mathrm{\Phi }(y)}{(yx)^2}},1/2x1/2,`$
where $`\mathrm{\Phi }`$ is a wave function of a bound state, $`M`$ denotes the normalized meson mass, $`m`$ stands for the normalized (anti-)quark mass and $`\mathrm{}`$ denotes the finite part integral. In Eq. (1), we shifted the variable $`x`$ total amount of $`1/2`$ compared with the variable in Refs. , in order to show the symmetry of the wave function transparently.
According to Harada and collaborators , one can expect that the wave function could be expanded as follows:
$`\mathrm{\Phi }(x)=\underset{N\mathrm{}}{lim}{\displaystyle \underset{j=0}{\overset{N}{}}}a_j(14x^2)^{\beta +j}.`$ (2)
The exponent $`\beta `$ and the normalized quark mass $`m`$ are related to each other by the equation
$`(m^21)+\beta \pi \mathrm{cot}\beta \pi =0.`$ (3)
The authors of references adopted the positive smallest solution $`\beta _0(m)`$ of Eq. (3) as $`\beta `$ in Eq. (2).
Mo and Perry, and Harada and his collaborators presented effective way to determine the coefficients $`a_j`$โs. We will briefly reproduce their procedures. By the use of the expansion in Eq. (2) truncated to given finite number $`N`$ for the wave function $`\mathrm{\Phi }`$, we multiply both sides of Eq. (1) by $`(14x^2)^{\beta +i}`$ and integrate them over x, then we obtain
$`M^2\widehat{N}\stackrel{}{a}=\widehat{H}\stackrel{}{a},\stackrel{}{a}={}_{}{}^{t}[a_0,a_1,\mathrm{},a_{n1}].`$ (4)
In order to obtain eigenvalues of the generalized eigenvalue equation, we have to solve the eigenvalue problem for norm matrix $`\widehat{N}`$, first, i.e.,
$`\widehat{N}\stackrel{}{v}_i=\lambda _i\stackrel{}{v}_i.`$ (5)
Next, we introduce a transformation matrix $`\widehat{W}`$ by
$`\widehat{W}=[{\displaystyle \frac{\stackrel{}{v}_1}{\stackrel{}{v}_1\sqrt{\lambda _1}}},\mathrm{},{\displaystyle \frac{\stackrel{}{v}_n}{\stackrel{}{v}_n\sqrt{\lambda _n}}}].`$ (6)
Then, we can transform Eq. (4) into a usual eigenvalue problem of the form
$`M^2\stackrel{}{b}={}_{}{}^{t}\widehat{W}\widehat{H}\widehat{W}\stackrel{}{b},\stackrel{}{a}=\widehat{W}\stackrel{}{b}.`$ (7)
For $`N=3`$ and $`m=0.01`$, we find, for the ground state boson,
Fig. 1
$`\beta `$ $`=`$ $`0.00552328,M^2=0.0366342,`$
$`a_0`$ $`=`$ $`1,a_1=0.00203562,a_2=0.000579369,a_3=0.000165813.`$ (8)
The values of the LHS and the RHS of Eq. (1) are shown in Fig. 1. In Fig. 1, the solid line represents the LHS and the dotted line stands for the RHS. The RHS with $`N=9`$, which is indicated by the dashed line, is exhibited for comparison. The coincidence of the LHS and the RHS is not inadmissible for small values of $`x`$. On the other hand, for $`x\pm 1/2`$, the difference between the LHS and the RHS is not allowable. Only the RHS has sharp spikes at the end points. This behavior is not changed much even if we improve the order of approximation.
We have to note here that in order to solve the generalized eigenvalue problem, the norm matrix should be positive definite. We cannot advance the above procedure beyond $`N12`$, because some of the eigenvalues of the norm matrix $`\widehat{N}`$ become almost zero or negative. These facts strongly suggest that the conventional basis function is not appropriate basis function for 2 dimensional field theories.
## 3 New Basis Function
In order to introduce new basis function, we first notice that there are infinite solutions $`\beta _1<\beta _2<\mathrm{}`$ of Eq. (3) in addition to solution $`\beta _0`$. We are easily led to
$`0\beta _n\beta _0n<1/2,`$
$`0<\beta _n\beta _k(nk)1,n>k.`$ (9)
The above relations imply that Eq. (2) never incorporates terms like $`(14x^2)^{\beta _n+j}`$ with positive integer $`n`$ and non-negative integer $`j`$.
We posit that the wave function is given by an infinite series
$`\mathrm{\Phi }(x)=\underset{N\mathrm{}}{lim}{\displaystyle \underset{n=0}{\overset{N}{}}}{\displaystyle \underset{j=0}{\overset{Nn}{}}}c_{n}^{}{}_{}{}^{j}\left(14x^2\right)^{\beta _n(m)+j}.`$ (10)
Substituting Eq. (10) into Eq. (1), we have
$`0`$ $`=`$ $`M^2\mathrm{\Phi }(x){\displaystyle _{1/2}^{1/2}}๐yH(x,y)\mathrm{\Phi }(y)|_{14x^2=ฯต}`$ (11)
$`=`$ $`4{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}c_{n}^{}{}_{}{}^{0}\left(m^21+\pi \beta _n\mathrm{cot}\pi \beta _n\right)ฯต^{\beta _n1}`$
$`+{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{j=0}{\overset{\mathrm{}}{}}}F_{nj}(m;\beta _n;M;c\mathrm{})ฯต^{\beta _n+j}`$
$`+{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}G_k(m;\beta _0,\beta _1\mathrm{};M;c\mathrm{})ฯต^k.`$
Of course, the first line in Eq. (11) cancels automatically because of the definition of $`\beta _n`$โs. Suppose that we truncate series in Eq. (10) to $`O(ฯต^{\beta _N})`$. That is, we set $`c_{n}^{}{}_{}{}^{j}=0`$ for $`n+j>N`$. We demand Eq. (11) to hold up to $`O(ฯต^{\beta _N1})`$. Then we have $`N(N+3)/2`$ non-trivial equations. On the other hand, there are $`(N+1)(N+2)/2`$ unknown parameters. The number of parameters is larger than that of non-trivial equations by 1. Thus, we can solve the equations for $`c_{n}^{}{}_{}{}^{j}`$ in terms of $`M^2`$. Another equation of use to us is obtained by multiplying both sides of Eq. (1) by $`\mathrm{\Phi }(x)`$ and integrating them over x,
$`M^2{\displaystyle _{1/2}^{1/2}}๐x\left|\mathrm{\Phi }(x)\right|^2`$ $`=`$ $`{\displaystyle _{1/2}^{1/2}}{\displaystyle _{1/2}^{1/2}}๐x๐y\mathrm{\Phi }(x)H(x,y)\mathrm{\Phi }(y).`$ (12)
For a given $`m`$, we put $`M^2=M_i^2`$. We can then solve Eq. (11) for $`c_{n}^{}{}_{}{}^{j}`$ in terms of $`M_i`$. We thus obtain the $`M_i`$ dependent truncated wave function, say,
$`\mathrm{\Phi }(x;M_i)`$. We can calculate a new mass eigenvalue $`M_{i+1}`$ using this wave function as
$`M_{i+1}^2={\displaystyle \frac{<\mathrm{\Phi }(M_i)|H|\mathrm{\Phi }(M_i)>}{<\mathrm{\Phi }(M_i)|\mathrm{\Phi }(M_i)>}}.`$ (13)
For $`N15`$ and $`m=0.01`$, mass $`M^2`$ converges in 5 iterations. For $`0<m<0.5`$, we obtain $`M^2`$โs and fit them by polynomials, as follows:
$`M^2(m)=3.6276m+3.5803m^2`$
$`+0.06836m^3+O(m^4).`$ (14)
It should be noted here that the
coefficients of $`m`$ are consistent with
Fig. 2
Bergknoffโs result. In order to see the efficacy of this new basis function expansion, we show the wave functions in Fig. 2. There, the thin solid line represents the LHS in Eq. (1), provided that the wave function was approximated by Eq. (10) with $`N=15`$. The dotted line denotes the RHS with $`N=2`$, the dot-dashed line exhibits the RHS with $`N=3`$, the dot-dot-dashed line represents the RHS with $`N=4`$, the dot-dash-dashed line stands for the RHS with $`N=5`$, and the dashed line exhibits the RHS with $`N=10`$. The thick solid line indicates the RHS in Eq. (1) with wave function given in Eq. (10) with $`N=15`$.
## 4 Summary and Discussion
In the previous section we have introduced the new basis function and calculated the mass eigenvalue of the bound state using the new basis function. We have found that (1) the new basis function gives an effective approximation of the wave function, and (2) the mass eigenvalues are consistent with the results of the precursors.
It should be noted that Eq.(10) is, mathematically, the most general expansion. This means that there is no room to introduce any other additional terms like $`d(14x^2)^\gamma `$ for $`\gamma \beta _n+j`$ with non-negative integers $`n`$ and $`j`$. If we introduce such terms, then the following equality should hold
$`0=4d\left(m^21+\pi \gamma \mathrm{cot}(\pi \gamma )\right)(14x^2)^{\gamma 1}.`$ (15)
This demands that $`d0`$.
## Acknowledgements
The author would like to thank Professor K. Tanaka and Professor G.J. Aubrecht for comments and discussions during the early stage of this work. He is also grateful to Dr. Harada for useful discussions. This work was partialy supported by the Grants-in-Aid for Scientific Research of Ministry of Education, Science and Culture of Japan (No. 10640198).
## References |
warning/0003/cond-mat0003437.html | ar5iv | text | # First-order quantum phase transition in the orthogonal-dimer spin chain
## I Introduction
Recent extensive experimental and theoretical investigations on low-dimensional spin systems with frustration have been providing a variety of interesting topics. Among others, the spin-gap compound found recently by Kageyama et al., $`\mathrm{SrCu}_2(\mathrm{BO}_3)_2`$, exhibits a number of nontrivial properties due to strong frustration. In particular, the discovery of the magnetization plateaus at 1/3, 1/4 and 1/8 of the full moment has been stimulating further intensive studies. The remarkable point claimed by Miyahara and Ueda is that this frustrated system is a prototypical example of the two-dimensional (2D) version of the orthogonal-dimer model, whose unique structure gives rise to various unusual properties. This 2D model is known to be equivalent to the frustrated Shastry-Sutherland model on a square lattice with some diagonal bonds. It has been recently shown that there exists a novel first-order quantum phase transition between the dimer and plaquette phases in the 2D orthogonal-dimer model, which is accompanied by the jump and cusp singularities in the spin gap near the transition point. What is most interesting is that the compound $`\mathrm{SrCu}_2(\mathrm{BO}_3)_2`$ may be located in the vicinity of the first-order transition point.
As seen from the above studies, characteristic properties in the 2D system around the first-order transition point are certainly caused by the strong frustration common to this class of the orthogonal-dimer spin systems. Therefore, to clarify the essential properties of the system, further systematic studies on the quantum phase transition are highly desired. The 1D version of the orthogonal-dimer model may be the most appropriate system for this purpose, because it is the simplest model which possesses the frustration effect due to the dimer-plaquette structure.
Motivated by these hot topics, in this paper, we study the first-order transition in the 1D orthogonal-dimer spin chain in detail. The Hamiltonian we shall deal with is,
$`=J{\displaystyle \underset{(i,j)}{}}๐_i๐_j+J^{}{\displaystyle \underset{<i,j>}{}}๐_i๐_jH{\displaystyle \underset{i}{}}S_i^z,`$ (1)
where $`๐_i`$ is the $`s=1/2`$ spin operator at the $`i`$-th site, and the indices $`(i,j)`$ and $`<i,j>`$ represent the summation over intra- and inter-dimer pairs, respectively (see Fig. 1). The magnetic field is denoted as $`H`$ for which we set $`g\mu _\mathrm{B}=1`$ for convenience. Both of the exchange couplings $`J`$ and $`J^{}`$ are assumed to be antiferromagnetic. We shall use the normalized parameters $`j=J^{}/J`$ and $`h=H/J`$ in the following discussions.
It was previously shown by Ivanov and Richter that the above orthogonal-dimer chain undergoes the first-order quantum phase transition between the dimer phase (for small $`j`$) and the plaquette phase (for large $`j`$) as the coupling ratio $`j`$ is varied. However, the unique properties inherent in this frustrated system have not been discussed in detail yet, especially around the critical point. In what follows, by means of the exact diagonalization (ED) and series expansion methods, we demonstrate that the excitation gap in this system exhibits nontrivial behaviors such as the jump and cusp singularities around the transition point, reflecting the dimer-plaquette dual properties characteristic of this frustrated system. We also point out that such properties are not specific to the 1D system but also common to the 2D orthogonal-dimer system. By computing the magnetization curve by means of the ED together with the density-matrix renormalization group (DMRG), we further reveal that the formation of the magnetic plateaus is dramatically affected by the dimer-plaquette dual properties, and the resulting magnetic phase diagram has a rich structure.
This paper is organized as follows. In Sec. II, we investigate how the characteristic properties inherent in the orthogonal-dimer chain emerge in the low-energy excitations around the first-order transition point by means of the ED, the DMRG and the series expansion method. In Sec. III, by calculating the magnetization curve by the ED and the DMRG, we show that the plateau formation in the magnetization changes its character around the transition point, reflecting the strong frustration effects. The last section is devoted to summary and discussions.
## II zero-field properties
In this section, we investigate the quantum phase transition in the orthogonal-dimer chain in the absence of the magnetic field, by exploiting the ED and the series expansion methods.
### A Ground state
We start with the ground state properties in the dimer phase. When $`j=0`$, the system is reduced to an assembly of the decoupled dimers denoted by the solid line in Fig. 1, for which the product of independent dimer-singlets gives the ground state. The remarkable point for the orthogonal-dimer chain is that this simple dimer-singlet state is always an exact eigenstate of the Hamiltonian (1) with the energy $`E_\mathrm{g}/JN=3/8`$ in the entire range of $`j`$ (see Fig. 2), where $`N`$ is the number of total sites. Accordingly, the dimer-singlet state should be the exact ground state up to a certain critical value of $`j`$.
In the opposite limit of large $`j`$, the ground state is given by the disordered singlet state which is adiabatically connected to the isolated plaquette-singlets denoted by the broken line in Fig. 1. Therefore, starting from the isolated plaquette singlets for $`j=\mathrm{}`$, we can evaluate the ground state energy $`E_\mathrm{g}`$ for the plaquette phase with finite $`j`$ by means of the series expansion. Performing the plaquette expansion up to the eleventh order in $`j^1`$ combined with the first-order inhomogeneous differential method, we have obtained the ground state energy $`E_\mathrm{g}`$ rather precisely for the plaquette phase. The result is shown as the solid line in Fig. 2. It is seen that the ground state energy of the plaquette state coincides with that of the dimer state at a certain value of $`j`$, at which the first-order quantum phase transition occurs. The critical value is estimated as $`j_c=0.81900`$, which is further confirmed to be accurate up to the above figure by the DMRG calculation for the infinite chain. We note here that the present results for the ground state are consistent with those obtained by Ivanov and Richter.
### B Spin excitations
Let us move to the spin excitation spectrum. In the dimer phase $`(0<j<j_c)`$, we can construct a low-energy triplet excitation by substituting a local triplet for one of singlet-dimers forming the ground state. The remarkable point is that the hopping of this local triplet across the singlet-dimer such as the $`1`$-$`3`$ dimer in Fig. 1 is forbidden by its characteristic crystal structure. This implies that an excited triplet state is completely localized inside the finite strip $`(N=6)`$, and has no dispersion for its spectrum. Thus we can exactly estimate the spin gap from the ED calculation for the cluster of $`N=6`$. In Fig. 3, the ED results are shown as the solid line in the dimer phase.
In the plaquette phase $`(j>j_c)`$, it is naively expected that the low-energy excitation is described in terms of a triplet excitation on the isolated plaquette. However, the frustrating diagonal interaction in each plaquette should considerably affect the excitation spectrum. To see this, it is instructive to examine the energy-level structure of the isolated plaquette with the diagonal bond (see Fig. 1), whose Hamiltonian reads
$`_{\mathrm{plaquette}}=J๐ฌ_1๐ฌ_3+J^{}\left(๐ฌ_1+๐ฌ_3\right)\left(๐ฌ_2+๐ฌ_4\right).`$ (2)
We list the energy eigenvalues of the plaquette in Table I, where $`S_{13}`$ ($`S_{24}`$) is a quantum number of the spin $`๐ฌ_1+๐ฌ_3`$ $`(๐ฌ_2+๐ฌ_4)`$ and $`S`$ is that for the total spin.
It is seen in this table that the eigenstates of the isolated plaquette are classified into two sectors: the $`D`$-sector for $`S_{13}=0`$ and the $`P`$-sector for $`S_{13}=1`$. For $`j>1/2`$, the ground state is always singlet in the $`P`$-sector, whereas both of two separated sectors appear for the excitations when the coupling ratio $`j`$ is varied. When $`1/2<j<1`$, the lowest excitation is given by the $`S_{24}`$-singlet and $`S_{24}`$-triplet in the $`D`$-sector, which are four-fold degenerate with the excitation energy $`\mathrm{\Delta }E/J=2j1`$. In the regime $`j>1`$, the first excitation is a $`P`$-sector triplet with the excitation energy $`\mathrm{\Delta }E/J=j`$. Note that as far as the singlet-triplet excitations in the $`P`$-sector are concerned, there are no contributions from the $`D`$-sector. This simple fact allows us to replace $`S_{13}`$ in the plaquette with the $`s=1`$ spin, which will be used below to perform the mixed-spin cluster expansion for the $`P`$-sector.
Keeping the above properties in mind, we now turn back to the orthogonal-dimer chain in the thermodynamic limit, by taking into account the interaction among plaquettes. Note that excitations in the $`D`$-sector are still completely separated from those in the $`P`$-sector, according to the constraint specific to the orthogonal-dimer structure. This remarkable fact enables us to study the excitation spectrum for the $`D`$-sector and the $`P`$-sector separately. Namely, we perform the plaquette expansion up to the eleventh order in $`j^1`$ for the $`D`$-sector, whereas for the $`P`$-sector we exploit the mixed-spin cluster expansion up to the seventh order in $`j^1`$ with a starting spin configuration $`1/211/2`$. Applying the first-order inhomogeneous differential method to the obtained series, we then deduce the lowest excitations both for the $`D`$-sector and the $`P`$-sector, as shown in Fig. 3. In this figure, the dashed line in the plaquette phase represents the energy for four-fold degenerate excitations in the $`D`$-sector whereas the solid line is for a triplet excitation in the $`P`$-sector. It is seen that the energy of two kinds of excitations intersects each other at $`j_c^{}=0.872`$, giving rise to a cusp singularity in the spin gap as a function of $`j`$. In the region $`j>j_c^{}`$, both of the ground state and the lowest excited state belong to the $`P`$-sector on each plaquette, as is the case for the simple plaquette chain. On the other hand, when $`j_c<j<j_c^{}`$, the lowest excitation is described by four-fold degenerate level in the $`D`$-sector although the ground state still belongs to the plaquette phase ($`P`$-sector). The present plaquette expansion further uncovers that the wave function of the localized $`D`$-sector excitation with no dispersion is spatially extended over a number of sites, which shows sharp contrast to the dimer phase where the wave function of the triplet excitation is completely localized at a given site. This may imply that the four-fold degenerate excitations characterized by the $`D`$-sector possess the intermediate properties between those typical for the dimer phase and the plaquette phase. As $`j`$ is further decreased, we encounter the first-order quantum phase transition, at which the ground-energy of the two phases coincide with each other, while the first excited states have still different energies. Therefore, the spin gap at the transition point jumps from $`\mathrm{\Delta }^+(j_c)=0.3590(2)`$ for the plaquette phase to $`\mathrm{\Delta }^{}(j_c)=0.32309`$ for the dimer phase.
In order to complement the series-expansion results, we also compute the lowest triplet excitation by the ED method for the finite chain ($`N=16,24`$) with the periodic boundary condition. The obtained spin gap $`\mathrm{\Delta }`$ is shown as the crosses and the open circles in Fig. 3. It is seen that they are in fairly good agreement with the results obtained by the series expansion method, except for the vicinity of the first-order transition point $`j_c`$, where the finite-size effect in the ED still remains. Here it should be noticed that the present results provide much more detailed information for spin excitations than those obtained by Ivanov and Richter, where the spin gap as a function of $`j`$ has only a cusp-like singularity at the critical point. We also note that the ladder system with a similar orthogonal-dimer structure has the jump and the cusp singularities in the spin gap.
## III Plateaus in the magnetization curve
In this section, we study the magnetic properties of the orthogonal-dimer chain. By using the ED and the DMRG methods, we calculate the magnetization curve as a function of the magnetic field $`h(=H/J)`$, and show that their characteristic behavior stems from the dimer-plaquette dual properties found for the excitations in the previous section. To clearly understand the magnetization curve in the low-field region, we shall use the description based on hard-core bosons, in which low-energy triplet excitations are regarded as hard-core bosons.
### A Dimer phase
Let us start with the dimer phase. In Fig. 4, we show the magnetization curve for $`j=0.7(<j_c)`$ obtained by the ED method for the finite chain of $`N=8,16`$ and $`24`$. We also show the results obtained by the DMRG method. In this figure, the plateaus appear in the magnetization curve at 1/4 and 1/2 of the full moment ($`m=1/2`$) clearly.
The mechanism of the $`1/2`$-plateau formation is simply understood according to the unit-cell structure of the chain. As will be shown below, the $`1/2`$-plateau appears in the entire range of $`j`$. On the other hand, the $`1/4`$-plateau may be realized by the commensurate state in which each triplet excitation is aligned periodically at every finite-strip of $`N=8`$, reflecting the completely-localized nature of the triplet excitation discussed in the previous section. Since the localized triplet does not correlate with those in the neighbor strips, the macroscopic number of the triplets become degenerate at the lower-critical field $`h_c`$ (= the spin gap $`\mathrm{\Delta }`$), where the magnetic first-order transition occurs. In Fig. 4, we can indeed see that the magnetization curve jumps from $`m=0`$ to $`m=1/8`$ (1/4-plateau) at $`h_c`$ without any finite-size correction. Increasing the magnetic field, the other first-order transition to the 1/2-plateau state occurs. It is seen in Fig. 4 that there still remains the large-finite size effect near this critical point, since the triplet excitations correlate with each other in contrast to the transition discussed above.
We here note that several different phases observed in the magnetization process are specified by the spatial arrangement of the spin quantum number, $`S_{13}`$, for the diagonal bond in each plaquette (see Table I). In the dimer phase, the $`D`$-sector ($`S_{13}`$=0) is realized in each plaquette and in the 1/4-plateau state, the $`D`$\- and the $`P`$-sectors are crystallized alternately. On the other hand, in the region of $`m>1/4`$, each diagonal bond in the plaquette forms $`S_{13}=1`$ (the $`P`$-sector), and hence the frustration does not play any role for the magnetization process in this region. Therefore by calculating the magnetization for the unfrustrated plaquette chain, we can simply reproduce that for the orthogonal-dimer chain in this region.
### B Plaquette phase
We recall that there are two sort of excitations characterized by the $`P`$\- and the $`D`$-sector in the plaquette phase, which may provide a variety of the magnetization process. More precisely, the hard-core boson description suggests three possible behaviors for the magnetization curves:
* case I: $`D`$-triplet excitations lie energetically lower than $`P`$-triplet excitations. The magnetization curve has a structure similar to that in the dimer-phase.
* case II: $`D`$-triplet excitations lie slightly above the bottom of the $`P`$-triplet dispersion curve (see Fig. 6). Thus both of the features specific to the $`D`$-triplet and the $`P`$-triplet excitations appear in the magnetization curve.
* case III: $`D`$-triplet excitations lie sufficiently higher above the spin gap for $`P`$-triplet excitations. The magnetization curve may be characterized solely by the $`P`$-triplet excitation.
Since the above classification is simply based on the hard-core boson description, the interaction among bosons becomes more important when the number of bosons (i.e. the strength of the external field) is increased, which requires more proper discussions beyond the hard-core boson description. In what follows, based on the numerical results, we confirm that the magnetization curves are indeed classified into the above three categories.
We start with the analysis of the case I. We show the magnetization process for $`j=0.86`$ in Fig. 5, where we can see the $`1/4`$\- and $`1/2`$-plateaus clearly. In particular the magnetization jumps from zero to the $`1/4`$ plateau at the critical field $`h_c`$ with a small finite-size correction.
In the plaquette phase, the 1/4-plateau is generated by the commensurate state in which the plaquette singlet and triplet in the $`D`$-sector are crystallized alternately. The characteristic point in the 1/4-plateau state is that $`D`$-triplet excitations interact with each other when they are condensed to form the $`1/4`$-plateau, in contrast to those in the dimer phase where each triplet excitation is localized completely. We can indeed see this interaction effect in the ED spectrum, where the triplets gain the condensation energy, by which the critical field $`h_c`$ shifts to a slightly lower field than the spin-gap value. We note here that since the mechanism of the 1/4-plateau formation (first-order transition) at $`h_c`$ is essentially the same in the dimer phase and in the plaquette phase, so that the critical field $`h_c`$ does not show the discontinuity at $`j=j_c`$ as a function of $`j`$ (see Fig.9), being contrasted to the discontinuity observed in the spin gap at $`j=j_c`$. Increasing the field, the magnetic first-order transition takes place to the state in which $`S_{13}=1`$ is formed in each plaquette. Beyond this critical field, the magnetization traces the curve which is the same as that for unfrustrated plaquette chain.
In the case II, the $`D`$-triplet excitation level lies slightly above the bottom of the $`P`$-triplet dispersion curve, as shown in Fig.6.
The coexistence of two-kinds of distinct excitations influences the magnetization curve considerably, as seen in Fig. 7. As $`h`$ increases beyond the critical field $`(h_c=\mathrm{\Delta })`$, the magnetization should develop with $`m(hh_c)^{1/2}`$, since it is dominated by the $`P`$-triplet excitation whose dispersion relation is quadratic near the bottom $`(q=\pi /2)`$. In fact, we can see a staircase structure below the $`1/4`$-plateau for the ED results in Fig. 7, implying that the jump is now changed to the continuous increase in the magnetization. With slightly increasing $`h`$, however, the magnetization may stop to increase continuously and jump to the $`1/4`$-plateau. Beyond the 1/4-plateau, we again encounter the first-order phase transition, which is accompanied by the jump in $`m`$. This jump singularity is followed by the continuous increase in $`m`$, as already discussed in the case I.
In the case III, the excitation energy of the $`D`$-triplet is pushed up in higher energy region of the dispersion curve. Thus, before the Zeeman energy lowers the $`D`$-triplet level down to zero, a number of bosons are accommodated in the $`P`$-triplet excited levels, and then the interaction effect between bosons becomes too significant to use the hard-core boson description. In this parameter region, we find that the $`1/4`$-plateau completely disappears and the magnetization curve becomes smooth up to the $`1/2`$-plateau.
Consequently, we end up with the magnetic phase diagram on the $`(jh)`$ plane as shown in Fig. 9.
In this figure, (a) and (aโ) indicate the dimer-singlet and plaquette-singlet phase, respectively. The region (b) exhibits the plateau in the magnetization curve at 1/4 of the full moment. We can see that the 1/4-plateau phase with the dimer-singlet structure exists not only in the dimer phase $`(j<j_c)`$ but also in a part of the plaquette phase. In the region (c), the $`1/2`$-plateau appears, which is stable in the whole parameter region of $`j`$. All spins are polarized in (d). The bold (solid) line shows the phase boundary of the first- (second-) order transition. It is now seen that the magnetic phase diagram possesses quite a rich structure, which is indeed caused by the dimer-plaquette dual properties inherent in this strongly frustrated spin chain. Here we wish to mention that the ladder system with a similar orthogonal-dimer structure has the jump and the cusp singularities in the magnetization curve.
## IV Summary and Discussions
We have studied the low-energy properties of the orthogonal-dimer spin chain with a frustrated dimer-plaquette structure, by means of the exact diagonalization, the density matrix renormalization group and the series expansion method. When the inter-dimer couplings are varied, the first-order quantum phase transition occurs at the critical value $`j_c=0.81900`$ between the dimer-singlet and plaquette-singlet phases. In the dimer phase, the ground state is exactly given by the decoupled dimer-singlet state, and triplet excitations over it are completely localized. In the plaquette phase, on the other hand, we have found that the excited states are characterized by the dual properties: the coexistence of plaquette-like and dimer-like excitations. As a result, both of the cusp and the jump appear in the spin gap as a function of $`j`$, which are caused by the level-crossing between the excited states. We have also clarified how the above dual properties influence the magnetization process of the chain by analyzing the numerical results in terms of the hard-core boson description. In particular, we have shown that three types of the magnetization curves appear around the magnetization plateau at $`1/4`$ of the full moment, according to how two kinds of triplet excitations change their relative positions energetically. We have thus found that the magnetic phase diagram on the $`(jh)`$ plane has a rich structure.
Although we have dealt with the one-dimensional chain in this paper, we believe that the present study captures some essential properties of the first-order phase transition and the resulting rich structure in low-lying excited states even in two dimension. For the compound $`\mathrm{SrCu}_2(\mathrm{BO}_3)_2`$, it is indeed pointed out that the low-energy triplet excitation has quite small mobility reflecting the orthogonal-dimer structure. Since this compound is located in the vicinity of the first-order phase transition point, the present results encourage us to interpret the plateau-formation mechanism of the 2D compound in terms of the dual properties originating from the dimer and plaquette structures. In particular, it is an interesting problem to study how the characteristic properties in 1D are changed into those in 2D by adiabatically introducing the inter-chain coupling, which may provide a clue to fully understand the physics in 2D orthogonal-dimer system.
## Acknowledgement
The work is partly supported by a Grant-in-Aid from the Ministry of Education, Science, Sports, and Culture. A. K. and K. O. are supported by the Japan Society for the Promotion of Science. A part of our computational program is based on the TITPACK ver. 2 by H. Nishimori. Numerical computations in this work was carried out at the Yukawa Institute Computer Facility. |
warning/0003/nucl-ex0003007.html | ar5iv | text | # Baryon Rapidity Loss in Relativistic Au+Au Collisions
## Abstract
An excitation function of proton rapidity distributions for different centralities is reported from AGS Experiment E917 for Au+Au collisions at 6, 8, and 10.8 GeV/nucleon. The rapidity distributions from peripheral collisions have a valley at mid-rapidity which smoothly change to distributions that display a broad peak at mid-rapidity for central collisions. The mean rapidity loss increases with increasing beam energy, whereas the fraction of protons consistent with isotropic emission from a stationary source at midrapidity decreases with increasing beam energy. The data suggest that the stopping is substantially less than complete at these energies.
PACS number(s): 25.75.-q, 13.85.Ni, 21.65.+f
Nuclear matter is believed to be compressed to high baryon density ($`\rho _B`$) during central collisions of heavy nuclei at relativistic energies. In the interaction region of the colliding nuclei, nucleons undergo collisions which reduce their original longitudinal momentum. This loss of rapidity is an important characteristic of the reaction mechanism, and is often referred to as stopping. The rapidity loss for beam nucleons has been extensively studied in p+A reactions and more recently in heavy-ion reactions. Relativistic heavy-ion collisions are unique in the sense that secondary collisions of excited baryons are expected to contribute to the rapidity loss leading to simultaneous stopping of many nucleons within the interaction volume. The $`\rho _B`$ thus reached may be large enough to induce phase transitions, such as quark deconfinement and/or chiral symmetry restoration . The observation of large numbers of baryons at mid-rapidity is indicative of compression to high $`\rho _B`$, but the quantitative connection to $`\rho _B`$ is only possible via model calculations. An important key to our understanding of this phenomenon is systematic data for the rapidity distributions of baryons over a broad range of conditions, such as beam energy and centrality.
In this Letter, we present an excitation function of the centrality dependence of proton rapidity distributions from Au+Au collisions at 6, 8, and 10.8 GeV/nucleon. At these energies, the measured ratio of antiprotons to protons is very small ($`<0.03\%`$) and the production of protons from $`\mathrm{\Lambda }`$ decay contributes less than 5% to the total yield. The measured protons can therefore be considered to directly reflect the distribution in rapidity of the initial baryons (assuming that the neutron rapidity distribution has a similar shape). The measured rapidity distributions at all three beam energies show a similar evolution with centrality. They are clearly bimodal in shape for peripheral collisions and change to shapes which for central collisions may still be bimodal in nature. This suggests that the degree of stopping in central collisions is not complete, even in this heavy system. In addition, it is shown that the protons do not end up as an isotropically emitting source at mid-rapidity, but retain a fair degree of their initial longitudinal motion. This phenomenon is represented by a parameterized fit to the data with a superposition of longitudinally moving sources.
Experiment E917 measured Au+Au reactions at beam kinetic energies of 6, 8, and 10.8 GeV/nucleon at the Brookhaven AGS, and was the final experiment in the series E802/E859/E866/E917. The experimental apparatus in E917 consisted of a series of beam-line detector arrays which were used for global event characterization, and a large rotatable magnetic spectrometer used to track and identify particles. The tracking system of the spectrometer consisted of a series of drift and multi-wire ionization chambers, which bracketed either side of a dipole magnet, followed by a segmented time-of-flight wall of vertical scintillator slats. The data presented here were taken with a trigger that required at least one track in the spectrometer. The centrality of an event can be selected through either of two approximately equivalent methods: (1) from the multiplicity measured by a large acceptance device called the New Multiplicity Array, or (2) from the energy deposited in the zero degree calorimeter. The data were sorted off-line into different centrality classes, where normalization was provided by prescaled interaction triggers. For each beam energy, the five event classes are reported as the percentage of the total interaction cross section ($`\sigma _{\mathrm{int}}`$ = 6.8 b), corresponding to (0-5)%, (5-12)%, (12-23)%, (23-39)%, and (39-81)% ((39-77)% for 10.8 GeV/nucleon) of $`\sigma _{\mathrm{int}}`$.
The systematic uncertainty on the normalization of the measured invariant spectra and rapidity distributions is dominated by the uncertainty of the single-track efficiency and the loss of tracks due to hit-blocking. The tracking uncertainty increases from 5% to 10% towards mid-rapidity. For peripheral collisions, there is a 10% uncertainty in the cross section, which decreases to 5% for central collisions. These uncertainties lead to a total systematic uncertainty of 15% independent of centrality, with a 5% relative uncertainty across beam energies.
The measured invariant yields of protons from Au+Au collisions at 8 GeV/nucleon are shown in ten rapidity intervals in Fig. 1 as a function of the transverse mass, $`m_t=\sqrt{p_t^2+m_0^2}`$. The spectra at 8 GeV/nucleon are representative of the quality of the data at other beam energies. The errors are statistical only. The data were fit with a Boltzmann-form function in $`m_t`$:
$`{\displaystyle \frac{1}{2\pi m_t}}{\displaystyle \frac{d^2N}{dm_tdy}}={\displaystyle \frac{dN/dy}{2\pi (m_0^2T+2m_0T^2+2T^3)}}m_te^{(m_tm_0)/T}`$ (1)
where $`T`$ and $`dN/dy`$ are free parameters. The data points were weighted according to their statistical errors. As has been previously reported, for Au+Au collisions at 10.8 GeV/nucleon, the proton spectra cannot be satisfactorily described by a single exponential function. In contrast, the above form reproduces the spectra well with $`\chi ^2`$ per d.o.f in the range 0.5 - 2.0, and provides the inverse slope parameter ($`T`$) and the rapidity density ($`dN/dy`$) for each rapidity interval.
The centrality dependence of the rapidity density at each beam energy is shown in Fig. 2. The data are shown relative to the rapidity of the center-of-mass of the system ($`y_{\mathrm{cm}}`$) which equals 1.35, 1.47, and 1.61 for beam kinetic energies of 6, 8, and 10.8 GeV/nucleon, respectively. The data points reflected about mid-rapidity are shown as open symbols. The errors were calculated from the fitting procedure. The present values of $`dN/dy`$ for the most central event class at 10.8 GeV/nucleon are in good agreement with the previously reported values from the E866 Collaboration.
For each beam energy, a common trend is observed in the evolution of the shape of the rapidity distributions as a function of centrality. For the most peripheral event class the distribution has a minimum value of $`dN/dy`$ 6 at mid-rapidity. This concave shape persists to the next most central event class, corresponding to a centrality cut of (23-39)%, consistent with the expectation that most of the participant protons reside at beam rapidities following these relatively peripheral collisions. For more central collisions, the rapidity distributions become progressively flatter, and begin to develop a broad maximum at mid-rapidity for all beam energies. The trend of the data suggests, however, that for the most central collisions, the distribution does not evolve to a single peak centered at mid-rapidity, but rather is consistent with two components each displaced from mid-rapidity, or with a set of sources spread throughout the rapidity range. To qualitatively emphasize this point, we have fit each of the measured distributions with two Gaussian peaks centered symmetrically about mid-rapidity. These are shown superimposed on the data for the three most central event classes at 8 GeV/nucleon in Fig. 3 and clearly display the trend. This result, somewhat surprisingly, implies that complete stopping is not achieved at AGS energies and that the longitudinal rapidity distribution is a result of transparency.
To quantify these results, we calculate the mean rapidity loss from the fits for the most central collisions $`\delta y=y_{beam}y`$ where $`y`$ refers to the Gaussians centered at positive rapidity. These values are listed in Table I. To compare with previous results for the most central data at 10.8 GeV/nucleon, we have also calculated the mean rapidity loss for protons in the restricted rapidity range $`1.11<yy_{\mathrm{cm}}<0`$ and find a value of 1.07$`\pm `$0.05, which is in good agreement with the value of 1.02$`\pm `$0.01 obtained for the (0-4)% most central Au+Au collisions at the same beam energy reported by Videbรฆk and Hansen .
The mean rapidity losses determined from our data show a systematic increase with incident energy. The mean rapidity loss does not, however, fully characterize all the features of the final rapidity distribution. Despite the fact that a considerable amount of the original longitudinal momentum of the incident particles has been lost, the situation is clearly far from one where they are completely stopped. Fig. 4 shows the $`dN/dy`$ distributions and inverse slopes for the (0-5)% most central event classes at all three energies plotted as solid points. The dashed curves in Fig. 4 show the expected distribution for complete stopping - i.e. isotropic emission from a source at rest in the center-of-mass system emitting protons with a Boltzmann energy distribution, the effective temperature of which is adjusted to reproduce the inverse slope of the transverse mass spectrum at mid-rapidity. The effective temperature thus accounts for the combined effects of radial expansion and temperature of this source.
Whereas this gives a good description of the rapidity dependence of the inverse slope parameters, it completely fails for the $`dN/dy`$ distributions. In fact, only a fraction of the observed yields can be accounted for in such a scenario. The yields can, however, be adequately accounted for by emission from a continuum of isotropic sources, uniformly distributed over a rapidity range $`y_{cm}\pm y_b`$. In order to account for the rapidity dependence of the inverse slope parameter it is, however, necessary to introduce a Gaussian rapidity dependency of the effective temperature $`T_{\mathrm{eff}}(y)=T_{\mathrm{eff}}^0\mathrm{exp}((yy_{\mathrm{cm}})^2/2\sigma _T^2)`$. The solid curves shown in Fig. 4 were obtained by a four parameter simultaneous fit of $`y_b;T_{\mathrm{eff}}^0;\sigma _T`$; and $`N_0`$ ($`dN/dy`$ normalization) to the experimental $`dN/dy`$ and inverse slope $`T(y)`$ values. Based on the fit $`dN/dy`$-distributions, the total number of protons within the range $`2<yy_{cm}<2`$ ($`_2^2\frac{dN}{dy}๐y`$) was found to be 155, 164, and 159 for $`E_{\mathrm{beam}}`$ = 6, 8, and 10.8 GeV/nucleon, respectively. Since the total initial number of protons (158) is approximately accounted for, this suggests that the extrapolation into the unmeasured region is quite reasonable.
The fraction of the observed protons corresponding to complete stopping ($`f_{iso}`$) was calculated as the ratio of the areas under the dashed and solid curves, respectively. The values of $`f_{iso}`$ at each energy are listed in Table I. The values of the absolute rapidity loss from the distributed source fits are also listed in Table I. They are in close agreement with the values from the Gaussian fits and show the same systematic increase with increasing incident energy. This result is different from observations in p+A collisions in which the rapidity loss was found to be independent of beam energy . This would suggest an increased role of secondary interactions which increase stopping in heavy-ion collisions. In contrast, the fractional rapidity loss, $`\delta y/\delta y_{max}`$, where $`\delta y_{max}=y_{beam}y_{cm}`$, is found to be essentially constant with beam energy from 6 to 158 GeV/nucleon, similar to the trend noted in Ref. .
In summary and conclusion, we have measured proton rapidity distributions for Au+Au collisions at three energies as a function of collision centrality. These distributions show a consistent evolution with increasing energy and centrality leading to maximal rapidity loss for the most central collisions and the highest energy. This result suggests the importance of secondary reactions in the heavy-ion case, contrary to the situation in p+A collisions. Nevertheless, the degree of stopping of the incident baryons is far from complete and only a fraction of the observed protons can be accounted for by the emission from a stopped isotropically emitting source. The remainder still possess considerable longitudinal momentum. It is not possible to say unequivocally from the present data alone whether or not this corresponds to a situation where the colliding nuclei are, to some extent, โtransparentโ, or a fully stopped and compressed system has re-expanded. However, the systematic behavior of the centrality dependence of the shapes of the rapidity distributions suggests, surprisingly, that the former is the case.
This work is supported by the U.S. Department of Energy under contracts with ANL (No. W-31-109-ENG-38), BNL (No. DE-AC02-98CH10886), MIT (No. DE-AC02-76ER03069), UC Riverside (No. DE-FG03-86ER40271), UIC (No. DE-FG02-94ER40865), and the University of Maryland (No. DE-FG02-93ER40802), the National Science Foundation under contract with the University of Rochester (No. PHY-9722606), and the Ministry of Education and KOSEF (No. 951-0202-032-2) in Korea. |
warning/0003/physics0003106.html | ar5iv | text | # 0 (E). FORWARD
## 0 (E). FORWARD
The energy quanta occured in 1900 in the work of Max Planck (Nobel prize, 1918) on the black body electromagnetic radiation. Planckโs โquanta of lightโ have been used by Einstein (Nobel prize, 1921) to explain the photoelectric effect, but the first โquantizationโ of a quantity having units of action (the angular momentum) belongs to Niels Bohr (Nobel Prize, 1922). This opened the road to the universalization of quanta, since the action is the basic functional to describe any type of motion. However, only in the 1920โs the formalism of quantum mechanics has been developed in a systematic manner. The remarkable works of that decade contributed in a decisive way to the rising of quantum mechanics at the level of fundamental theory of the universe, with successful technological applications. Moreover, it is quite probable that many of the cosmological misteries may be disentangled by means of various quantization procedures of the gravitational field, advancing our understanding of the origins of the universe. On the other hand, in recent years, there is a strong surge of activity in the information aspect of quantum mechanics. This aspect, which was generally ignored in the past, aims at a very attractive โquantum computerโ technology.
At the philosophical level, the famous paradoxes of quantum mechanics, which are perfect examples of the difficulties of โquantumโ thinking, are actively pursued ever since they have been first posed. Perhaps the most famous of them is the EPR paradox (Einstein, Podolsky, Rosen, 1935) on the existence of elements of physical reality, or in EPR words: โIf, without in any way disturbing a system, we can predict with certainty (i.e., with probability equal to unity) the value of a physical quantity, then there exists an element of physical reality corresponding to this physical quantity.โ Another famous paradox is that of Schrรถdingerโs cat which is related to the fundamental quantum property of entanglement and the way we understand and detect it. What one should emphasize is that all these delicate points are the sourse of many interesting and innovative experiments (such as the so-called โteleportationโ of quantum states) pushing up the technology.
Here, I present eight elementary topics in nonrelativistic quantum mechanics from a course in Spanish (โcastellanoโ) on quantum mechanics that I taught in the Instituto de Fรญsica, Universidad de Guanajuato (IFUG), Leรณn, Mexico, during the semesters of 1998.
Haret C. Rosu
## 0 (R). CUVรNT รNAINTE
Cuantele de energie au apฤrut รฎn 1900 ca o consecinลฃฤ a lucrฤrilor lui Max Planck (premiul Nobel 1918) asupra problemei radiaลฃiei de corp negru. โCuantele de luminฤโ planckiene au fost folosite de cฤtre Albert Einstein (premiul Nobel 1921) pentru a explica efectul fotoelectric, dar prima โcuantificareโ a unei mฤrimi cu unitฤลฃi de acลฃiune (momentul cinetic) se datoreazฤ lui Niels Bohr (premiul Nobel 1922). Aceasta a deschis drumul universalitฤลฃii cuantelor pentru cฤ acลฃiunea este funcลฃionala fundamentalฤ pentru descrierea oricฤrui tip de miลcare. Chiar ลi รฎn aceste condiลฃii, numai anii 1920 se considerฤ ca adevฤratul รฎnceput pentru formalismul cuantic, care a fost capabil sฤ ridice mecanica cuanticฤ la nivelul unei teorii fundamentale a universului ลi sฤ o transforme รฎntr-o sursฤ de numeroase succese tehnologice. Este foarte posibil ca multe dintre misteriile cosmologice sฤ se ascundฤ รฎn spatele diferitelor proceduri de cuantificare ale cรฎmpului nelinear gravitaลฃional ลi eventualele progrese รฎn aceastฤ direcลฃie ar putea contribui la o mai bunฤ รฎnลฃelegere a istoriei ลi evoluลฃiei universului. Pe de alta parte, aspectul informatic al mecanicii cuantice, care nu a fost mult investigat รฎn trecut, cunoaลte รฎn prezent o perioadฤ explozivฤ de cercetฤri รฎn ideea construirii aลa-numitelor โcalculatoare cuanticeโ.
รn domeniul filosofic este de menลฃionat cฤ รฎn mecanica cuanticฤ existฤ paradoxuri faimoase, care รฎncฤ se menลฃin รฎn polemicฤ ลi care reflectฤ dificultฤลฃile de logicฤ pe care le creazฤ modul de โgรฎndire cuanticฤโ (sau probabilisticฤ cuanticฤ ). Unul dintre cele mai cunoscute paradoxuri este cel al lui Einstein (care nu a acceptatรฎn mod total mecanica cuanticฤ), Podolsky ลi Rosen (EPR, 1935) รฎn legฤturฤ cu problema dacฤ existฤ sau nu โelemente adevฤrate de realitate fizicฤโ รฎn microcosmosul studiat cu metode cuantice (dupฤ Einstein, mecanica cuanticฤ interzice existenลฃa independentฤ a actului de mฤsurare de sistemele fizice mฤsurate). Alt paradox, de acelaลi rang de celebritate, este al โpisicii lui Schrรถdingerโ. Ceea ce trebuie subliniat รฎn legฤturฤ cu toate aceste puncte teoretice ลi metateoretice delicate este cฤ ele genereazฤ experimente foarte interesante (cum ar fi, de exemplu, cele referitoare la aลa-numita โteleportareโ ale stฤrilor cuantice) care impulseazฤ dezvoltarea tehnologicฤฤeea ce urmeazฤ sunt cรฎteva teme introductive รฎn mecanica cuanticฤ nerelativistฤ care au servit ca bazฤ pentru cursul de graduare รฎn mecanica cuanticฤ pe care l-am predat รฎn Institutul de Fizicฤ al Universitฤลฃii Statale Guanajuato din Mexic รฎn 1998.
Haret C. Rosu
## 1. POSTULATE CUANTICE
Urmฤtoarele 6 postulate se pot considera ca bazฤ pentru teorie ลi experiment รฎn mecanica cuanticฤ รฎn varianta sa cea mai folositฤ (standard).
1. Fiecฤ rei mฤrimi fizice โbine definitฤ clasicโ L รฎi corespunde un operator hermitic $`\widehat{L}`$.
1. Fiecฤrei stฤri fizice staลฃionare รฎn care se poate gฤsi un sistem fizic cuantic รฎi corespunde o funcลฃie de undฤ normalizatฤ $`\psi `$ ($`\psi _^2^2=1`$).
1. Mฤrimea fizicฤ L poate sฤ โiaโ experimental numai valorile proprii ale $`\widehat{L}`$. De aceea, valorile proprii trebuie sฤ fie reale, ceea ce are loc pentru operatori hermitici.
1. Rezultatul unei mฤsurฤtori pentru determinarea mฤrimii L este รฎntotdeauna valoarea medie $`\overline{L}`$ a operatorului $`\widehat{L}`$ รฎn starea $`\psi _n`$, care รฎn teorie este elementul de matrice diagonal
$`<\psi _n\widehat{L}\psi _n>=\overline{L}`$.
1. Elementele de matrice ale operatorilor coordonatฤ ลi moment carteziene $`\widehat{x_i}`$ ลi $`\widehat{p_k}`$, calculate intre funcลฃiile de undฤ f ลi g satisfac ecuaลฃiile de miลcare Hamilton din mecanica clasica รฎn forma:
$$\frac{d}{dt}<f\widehat{p_i}g>=<f\frac{\widehat{H}}{\widehat{x_i}}g>$$
$$\frac{d}{dt}<f\widehat{x_i}g>=<f\frac{\widehat{H}}{\widehat{p_i}}g>,$$
unde $`\widehat{H}`$ este operatorul hamiltonian, iar derivatele รฎn raport cu operatori se definesc รฎn punctul 3 al acestui capitol.
1. Operatorii $`\widehat{p_i}`$ ลi $`\widehat{x_k}`$ au urmฤtorii comutatori:
$$[\widehat{p_i},\widehat{x_k}]=i\mathrm{}\delta _{ik},$$
$$[\widehat{p_i},\widehat{p_k}]=0,$$
$$[\widehat{x_i},\widehat{x_k}]=0$$
$`\mathrm{}=h/2\pi =1.0546\times 10^{27}`$ erg.sec.
1. Corespondenลฃa cu o mฤrime fizicฤ L care are analog clasic $`L(x_i,p_k)`$
Aceasta se face substituind $`x_i`$, $`p_k`$ cu $`\widehat{x_i}`$ $`\widehat{p_k}`$. Funcลฃia L se presupune cฤ se poate dezvolta รฎn serie de puteri pentru orice valoare a argumentelor sale, adicฤ este analiticฤ . Dacฤ funcลฃia nu conลฃine produse $`x_kp_k`$, operatorul $`\widehat{L}`$ este hermitic รฎn mod automat.
Exemplu:
$`T=(_i^3p_i^2)/2m`$ $``$ $`\widehat{T}=(_i^3\widehat{p}^2)/2m`$.
Dacฤ L conลฃine produse mixte $`x_ip_i`$ ลi puteri ale acestora, $`\widehat{L}`$ nu este hermitic, รฎn care caz L se substituie cu $`\widehat{\mathrm{\Lambda }}`$, partea hermiticฤ a lui $`\widehat{L}`$ ($`\widehat{\mathrm{\Lambda }}`$ este un operator autoadjunct).
Exemplu:
$`w(x_i,p_i)=_ip_ix_i`$ $``$ $`\widehat{w}=1/2_i^3(\widehat{p_i}\widehat{x_i}+\widehat{x_i}\widehat{p_i})`$.
Rezultฤ deasemenea cฤ timpul nu este un operator fiind doar un parametru (care se poate introduce รฎn multe feluri). Aceasta pentru cฤ timpul nu depinde de variabilele canonice ci din contrฤ .
1. Probabilitate รฎn spectrul discret ลi continuu
Dacฤ $`\psi _n`$ este funcลฃie proprie a operatorului $`\widehat{M}`$, atunci:
$`\overline{L}=<n\widehat{L}n>=<n\lambda _nn>=\lambda _n<nn>=\delta _{nn}\lambda _n=\lambda _n`$.
Deasemenea, se poate demonstra cฤ $`\overline{L}^k=(\lambda _n)^k`$.
Dacฤ funcลฃia $`\varphi `$ nu este funcลฃie proprie a lui $`\widehat{L}`$ se foloseลte dezvoltarea รฎn sistem complet de f.p. ale lui $`\widehat{L}`$ ลi deci:
$`\widehat{L}\psi _n=\lambda _n\psi _n`$, $`\varphi =_na_n\psi _n`$
combinรฎnd aceste douฤ relaลฃii obลฃinem:
$`\widehat{L}\varphi =_n\lambda _na_n\psi _n`$.
Putem astfel calcula elementele de matrice ale operatorului L:
$`<\varphi \widehat{L}\varphi >=_{n,m}a_m^{}a_n\lambda _n<mn>=_ma_m^2\lambda _m`$,
ceea ce ne spune cฤ rezultatul experimentului este $`\lambda _m`$ cu o probabilitate $`a_m^2`$.
Dacฤ spectrul este discret: de acord cu P4 รฎnseamnฤ cฤ $`a_m^2`$, deci coeficienลฃii dezvoltฤrii รฎntr-un sistem complet de f.p., determinฤ probabilitฤtile de observare a valorii proprii $`\lambda _n`$.
Dacฤ spectrul este continuu: folosind urmฤtoarea definiลฃie
$`\varphi (\tau )=a(\lambda )\psi (\tau ,\lambda )๐\lambda `$,
se calculeazฤ elementele de matrice pentru spectrul continuu
$`<\varphi \widehat{L}\varphi >`$
$`=๐\tau a^{}(\lambda )\psi ^{}(\tau ,\lambda )๐\lambda \mu a(\mu )\psi (\tau ,\mu )๐\mu `$
$`=a^{}a(\mu )\mu \psi ^{}(\tau ,\lambda )\psi (tau,\mu )๐\lambda ๐\mu ๐\tau `$
$`=a^{}(\lambda )a(\mu )\mu \delta (\lambda \mu )๐\lambda ๐\mu `$
$`=a^{}(\lambda )a(\lambda )\lambda ๐\lambda `$
$`=a(\lambda )^2\lambda ๐\lambda `$.
รn cazul continuu se spune cฤ $`a(\lambda )^2`$ este densitatea de probabilitate de a observa v.p. $`\lambda `$ din spectrul continuu. Deasemenea, se satisface
$`\overline{L}=<\varphi \widehat{L}\varphi >`$.
Este comun sฤ se spunฤ cฤ $`<\mu \mathrm{\Phi }>`$ este (reprezentarea lui) $`\mathrm{\Phi }>`$ รฎn reprezentarea $`\mu `$, unde $`\mu >`$ este un vector propriu al lui $`\widehat{M}`$.
1. Definiลฃia unei derivate รฎn raport cu un operator
$`\frac{F(\widehat{L})}{\widehat{L}}=\mathrm{lim}_ฯต\mathrm{}\frac{F(\widehat{L}+ฯต\widehat{I})F(\widehat{L})}{ฯต}.`$
1. Operatorii de impuls cartezian
Care este forma concretฤ a lui $`\widehat{p_1}`$, $`\widehat{p_2}`$ y $`\widehat{p_3}`$, dacฤ argumentele funcลฃiilor de undฤ sunt coordonatele carteziene $`x_i`$ ?
Vom considera urmฤtorul comutator:
$`[\widehat{p_i},\widehat{x_i}^2]=\widehat{p_i}\widehat{x_i}^2\widehat{x_i}^2\widehat{p_i}`$
$`=\widehat{p_i}\widehat{x_i}\widehat{x_i}\widehat{x_i}\widehat{p_i}\widehat{x_i}+\widehat{x_i}\widehat{p_i}\widehat{x_i}\widehat{x_i}\widehat{x_i}\widehat{p_i}`$
$`=(\widehat{p_i}\widehat{x_i}\widehat{x_i}\widehat{p_i})\widehat{x_i}+\widehat{x_i}(\widehat{p_i}\widehat{x_i}\widehat{x_i}\widehat{p_i})`$
$`=[\widehat{p_i},\widehat{x_i}]\widehat{x_i}+\widehat{x_i}[\widehat{p_i},\widehat{x_i}]`$
$`=i\mathrm{}\widehat{x_i}i\mathrm{}\widehat{x_i}=2i\mathrm{}\widehat{x_i}`$.
รn general, se satisfac:
$`\widehat{p_i}\widehat{x_i}^n\widehat{x_i}^n\widehat{p_i}=ni\mathrm{}\widehat{x_i}^{n1}.`$
Atunci, pentru toate funcลฃiile analitice avem:
$`\widehat{p_i}\psi (x)\psi (x)\widehat{p_i}=i\mathrm{}\frac{\psi }{x_i}`$.
Acum, fie $`\widehat{p_i}\varphi =f(x_1,x_2,x_3)`$ modul รฎn care acลฃioneazฤ $`\widehat{p_i}`$ asupra lui $`\varphi (x_1,x_2,x_3)=1`$. Atunci:
$`\widehat{p_i}\psi =i\mathrm{}\frac{\psi }{x_1}+f_1\psi `$ ลi existฤ relaลฃii analoage pentru $`x_2`$ y $`x_3`$.
Din comutatorul $`[\widehat{p_i},\widehat{p_k}]=0`$ se obลฃine $`\times \stackrel{}{f}=0`$, ลi deci,
$`f_i=_iF`$.
Forma cea mai generalฤ a lui $`\widehat{p_i}`$ este $`\widehat{p_i}=i\mathrm{}\frac{}{x_i}+\frac{F}{x_i}`$, unde F este orice funcลฃie. Funcลฃia F se poate elimina folosind o transformare unitarฤ $`\widehat{U}^{}=\mathrm{exp}(\frac{i}{\mathrm{}}F)`$.
$`\widehat{p_i}=\widehat{U}^{}(i\mathrm{}\frac{}{x_i}+\frac{F}{x_i})\widehat{U}`$
$`=\mathrm{exp}^{\frac{i}{\mathrm{}}F}(i\mathrm{}\frac{}{x_i}+\frac{F}{x_i})\mathrm{exp}^{\frac{i}{\mathrm{}}F}`$
$`=i\mathrm{}\frac{}{x_i}`$
rezultรฎnd $`\widehat{p_i}=i\mathrm{}\frac{}{x_i}`$ $``$ $`\widehat{p}=i\mathrm{}`$.
1. Calculul constantei de normalizare
Orice funcลฃie de undฤ $`\psi (x)`$ $``$ $`^2`$ de variabilฤ $`x`$ se poate scrie รฎn forma:
$`\psi (x)=\delta (x\xi )\psi (\xi )๐\xi `$
ลi se poate considera aceastฤ expresie ca dezvoltare a lui $`\psi `$ รฎn f.p. ale operatorului coordonatฤ $`\widehat{x}\delta (x\xi )=\xi (x\xi )`$. Atunci, $`\psi (x)^2`$ este densitatea de probabilitate a coordonatei รฎn starea $`\psi (x)`$. De aici rezultฤ interpretarea normei
$`\psi (x)^2=\psi (x)^2๐x=1`$.
Intuitiv, aceastฤ relaลฃie ne spune cฤ sistemul descris de cฤtre funcลฃia $`\psi (x)`$ trebuie sฤ se gฤseascฤ รฎntr-un โlocโ pe axa realฤ, chiar dacฤ vom ลti doar aproximativ unde.
Funcลฃiile proprii ale operatorului impuls sunt:
$`i\mathrm{}\frac{\psi }{x_i}=p_i\psi `$, integrรฎnd se obลฃine $`\psi (x_i)=A\mathrm{exp}^{\frac{i}{\mathrm{}}p_ix_i}`$, $`x`$ ลi $`p`$ au spectru continuu ลi deci normalizarea se face cu โfuncลฃia deltaโ.
Cum se obลฃine constanta de normalizare ?
Se poate obลฃine utilizรฎnd urmฤtoarele transformฤri Fourier:
$`f(k)=g(x)\mathrm{exp}^{ikx}dx`$, $`g(x)=\frac{1}{2\pi }f(k)\mathrm{exp}^{ikx}dk.`$
Deasemenea se obลฃine cu urmฤtoarea procedurฤ :
Fie funcลฃia de undฤ nenormalizatฤ a particulei libere
$`\varphi _p(x)=A\mathrm{exp}^{\frac{ipx}{\mathrm{}}}`$ ลi formula
$`\delta (xx^{^{}})=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}\mathrm{exp}^{ik(xx^{^{}})}dx`$
Se vede cฤ
$`_{\mathrm{}}^{\mathrm{}}\varphi _p^{^{}}^{}(x)\varphi _p(x)๐x`$
$`=_{\mathrm{}}^{\mathrm{}}A^{}\mathrm{exp}^{\frac{ip^{^{}}x}{\mathrm{}}}A\mathrm{exp}^{\frac{ipx}{\mathrm{}}}dx`$
$`=_{\mathrm{}}^{\mathrm{}}A^2\mathrm{exp}^{\frac{ix(pp^{^{}})}{\mathrm{}}}dx`$
$`=A^2\mathrm{}_{\mathrm{}}^{\mathrm{}}\mathrm{exp}^{\frac{ix(pp^{^{}})}{\mathrm{}}}d\frac{x}{\mathrm{}}`$
$`=2\pi \mathrm{}A^2\delta (pp^{^{}})`$
ลi deci constanta de normalizare este:
$`A=\frac{1}{\sqrt{2\pi \mathrm{}}}`$.
Deasemenea rezultฤ cฤ f.p. ale operatorului impuls formeazฤ un sistem complet (รฎn sensul cazului continuu) pentru funcลฃiile de clasฤ $`^2`$.
$`\psi (x)=\frac{1}{\sqrt{2\pi \mathrm{}}}a(p)\mathrm{exp}^{\frac{ipx}{\mathrm{}}}dp`$
$`a(p)=\frac{1}{\sqrt{2\pi \mathrm{}}}\psi (x)\mathrm{exp}^{\frac{ipx}{\mathrm{}}}dx`$.
Aceste formule stabilesc legฤtura รฎntre reprezentฤrile x ลi p.
1. Reprezentarea p (de impuls)
Forma explicitฤ a operatorilor $`\widehat{p_i}`$ ลi $`\widehat{x_k}`$ se poate obลฃine din relaลฃiile de comutare, dar ลi folosind nucleele
$`x(p,\beta )=U^{}xU=\frac{1}{2\pi \mathrm{}}\mathrm{exp}^{\frac{ipx}{\mathrm{}}}x\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}}dx`$
$`=\frac{1}{2\pi \mathrm{}}\mathrm{exp}^{\frac{ipx}{\mathrm{}}}(i\mathrm{}\frac{}{\beta }\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}})`$.
Integrala are forma urmฤtoare: $`M(\lambda ,\lambda ^{^{}})=U^{}(\lambda ,x)\widehat{M}U(\lambda ^{^{}},x)๐x`$, ลi folosind $`\widehat{x}f=x(x,\xi )f(\xi )๐\xi `$, acลฃiunea lui $`\widehat{x}`$ asupra lui $`a(p)`$ $``$ $`^2`$ este:
$`\widehat{x}a(p)=x(p,\beta )a(\beta )๐\beta `$
$`=(\frac{1}{2\pi \mathrm{}}\mathrm{exp}^{\frac{ipx}{\mathrm{}}}(i\mathrm{}\frac{}{\beta }\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}})๐x)a(\beta )๐\beta `$
$`=\frac{i}{2\pi }\mathrm{exp}^{\frac{ipx}{\mathrm{}}}\frac{}{\beta }\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}}a(\beta )๐x๐\beta `$
$`=\frac{i\mathrm{}}{2\pi }\mathrm{exp}^{\frac{ipx}{\mathrm{}}}\frac{}{\beta }\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}}a(\beta )๐\frac{x}{\mathrm{}}๐\beta `$
$`=\frac{i\mathrm{}}{2\pi }\mathrm{exp}^{\frac{ix(\beta p)}{\mathrm{}}}\frac{}{\beta }a(\beta )๐\frac{x}{\mathrm{}}๐\beta `$
$`=i\mathrm{}\frac{a(p)}{\beta }\delta (\beta p)๐\beta =i\mathrm{}\frac{a(p)}{p}`$,
unde $`\delta (\beta p)=\frac{1}{2\pi }\mathrm{exp}^{\frac{ix(\beta p)}{\mathrm{}}}d\frac{x}{\mathrm{}}`$.
Operatorul impuls รฎn reprezentarea p se caracterizeazฤ prin nucleul:
$`p(p,\beta )=\widehat{U}^{}p\widehat{U}`$
$`=\frac{1}{2\pi \mathrm{}}\mathrm{exp}^{\frac{ipx}{\mathrm{}}}(i\mathrm{}\frac{}{x})\mathrm{exp}^{\frac{i\beta x}{\mathrm{}}}dx`$
$`=\frac{1}{2\pi \mathrm{}}\mathrm{exp}^{\frac{ipx}{\mathrm{}}}\beta \mathrm{exp}^{\frac{i\beta x}{\mathrm{}}}dx=\beta \lambda (p\beta )`$
rezultรฎnd $`\widehat{p}a(p)=pa(p)`$.
Ceea ce se รฎntรฎmplฤ cu $`\widehat{x}`$ ลi $`\widehat{p}`$ este cฤ deลi sunt operatori hermitici pentru toate f(x) $``$ $`L^2`$ nu sunt hermitici exact pentru funcลฃiile lor proprii.
Dacฤ $`\widehat{p}a(p)=p_oa(p)`$ ลi $`\widehat{x}=\widehat{x}^{}`$ $`\widehat{p}=\widehat{p}^{}`$, atunci
$`<a\widehat{p}\widehat{x}a><a\widehat{x}\widehat{p}a>=i\mathrm{}<aa>`$
$`p_o[<a\widehat{x}a><a\widehat{x}a>]=i\mathrm{}<aa>`$
$`p_o[<a\widehat{x}a><a\widehat{x}a>]=0`$
Partea stรฎngฤ este zero, รฎn timp ce dreapta este indefinitฤ , ceea ce este o contradicลฃie.
1. Reprezentฤrile Schrรถdinger ลi Heisenberg
Ecuaลฃiile de miลcare date prin P5 au diferite interpretฤri, datoritฤ faptului cฤ รฎn expresia $`\frac{d}{dt}<f\widehat{L}f>`$ se poate considera dependenลฃa temporalฤ ca aparลฃinรฎnd fie funcลฃiilor de undฤ fie operatorilor, fie atรฎt funcลฃiilor de undฤ cรฎt ลi operatorilor. Vom considera numai primele douฤ cazuri.
* Pentru un operator ce depinde de timp $`\widehat{O}=\widehat{O(t)}`$ avem:
$`\widehat{p_i}=\frac{\widehat{H}}{\widehat{x_i}}`$, $`\widehat{x_i}=\frac{\widehat{H}}{\widehat{p_i}}`$
$`[\widehat{p},f]=\widehat{p}ff\widehat{p}=i\mathrm{}\frac{f}{\widehat{x_i}}`$
$`[\widehat{x},f]=\widehat{x}ff\widehat{x}=i\mathrm{}\frac{f}{\widehat{p_i}}`$
ลi se obลฃin ecuaลฃiile de miลcare Heisenberg:
$`\widehat{p_i}=\frac{i}{\mathrm{}}[\widehat{p},\widehat{H}]`$, $`\widehat{x_i}=\frac{i}{\mathrm{}}[\widehat{x},\widehat{H}]`$.
* Dacฤ funcลฃiile de undฤ depind de timp, รฎncฤ se poate folosi $`\widehat{p_i}=\frac{i}{\mathrm{}}[\widehat{p_i},\widehat{H}]`$, pentru cฤ este o consecinลฃฤ numai a relaลฃiilor de comutare care nu depind de reprezentare
$`\frac{d}{dt}<f\widehat{p_i}g>=\frac{i}{\mathrm{}}<f[\widehat{p},\widehat{H}]g>`$.
Dacฤ acum $`\widehat{p_i}`$ ลi $`\widehat{H}`$ nu depind de timp ลi se ลฃine cont de hermiticitate se obลฃine:
$`(\frac{f}{t},\widehat{p_i}g)+(\widehat{p_i}f,\frac{g}{t})`$
$`=\frac{i}{\mathrm{}}(f,\widehat{p_i}\widehat{H}g)+\frac{i}{\mathrm{}}(f,\widehat{H}\widehat{p_i}g)`$
$`=\frac{i}{\mathrm{}}(\widehat{p}f,\widehat{H}g)+\frac{i}{\mathrm{}}(\widehat{H}f,\widehat{p_i}g)`$
$`(\frac{f}{t}+\frac{i}{\mathrm{}}\widehat{H}f,\widehat{p_i}g)+(\widehat{p_i}f,\frac{g}{t}\frac{i}{\mathrm{}}\widehat{H}g)=0`$
Ultima relaลฃie se รฎndeplineลte pentru orice pereche de funcลฃii $`f(x)`$ ลi $`g(x)`$ รฎn momentul iniลฃial dacฤ fiecare satisface ecuaลฃia
$`i\mathrm{}\frac{\psi }{t}=H\psi `$.
Aceasta este ecuaลฃia Schrรถdinger ลi descrierea sistemului cu ajutorul operatorilor independenลฃi de timp se cunoaลte ca reprezentarea Schrรถdinger.
รn ambele reprezentฤri evoluลฃia temporalฤ a sistemului se caracterizeazฤ prin operatorul $`\widehat{H}`$, care se obลฃine din funcลฃia Hamilton din mecanica clasicฤ .
Exemplu: $`\widehat{H}`$ pentru o particulฤ รฎn potenลฃial $`U(x_1,x_2,x_3)`$ este:
$`\widehat{H}=\frac{\widehat{p^2}}{2m}+U(x_1,x_2,x_3)`$, ลi รฎn reprezentarea x este:
$`\widehat{H}=\frac{\mathrm{}^2}{2m}_x^2+U(x_1,x_2,x_3)`$.
1. Legฤtura รฎntre reprezentฤrile S ลi H
P5 este corect รฎn reprezentฤrile Schrรถdinger ลi Heisenberg. De aceea, valoarea medie a oricฤrei observabile coincide รฎn cele douฤ reprezentฤri, ลi deci, existฤ o transformare unitarฤ cu care se poate trece de la o reprezentare la alta. O astfel de transformare este de forma $`\widehat{s}^{}=\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}}`$. Pentru a trece la reprezentarea Schrรถdinger trebuie folositฤ transformata Heisenberg $`\psi =\widehat{s^{}}f`$ cu $`f`$ ลi $`\widehat{L}`$, ลi pentru a trece la reprezentarea Heisenberg se foloseลte transformarea Schrรถdinger $`\widehat{\mathrm{\Lambda }}=\widehat{s^{}}\widehat{L}\widehat{s}`$ cu $`\psi `$ ลi $`\widehat{\mathrm{\Lambda }}`$. Se poate obลฃine ecuaลฃia Schrรถdinger dupฤ cum urmeazฤ : cum รฎn transformarea $`\psi =\widehat{s^{}}f`$ funcลฃia $`f`$ nu depinde de timp, vom deriva transformarea รฎn raport cu timpul pentru a obลฃine:
$`\frac{\psi }{t}=\frac{s^{}}{t}f=\frac{}{t}(\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}})f=\frac{i}{\mathrm{}}\widehat{H}\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}}f=\frac{i}{\mathrm{}}\widehat{H}\widehat{s^{}}f=\frac{i}{\mathrm{}}\widehat{H}\psi `$.
ลi deci, avem:
$`i\mathrm{}\frac{\psi }{t}=\widehat{H}\psi `$.
รn continuare obลฃinem ecuaลฃia Heisenberg: punรฎnd transformarea Schrรถdinger รฎn urmฤtoarea formฤ $`\widehat{s}\widehat{\mathrm{\Lambda }}\widehat{s^{}}=\widehat{L}`$ ลi derivรฎnd รฎn raport cu timpul se obลฃine ecuaลฃia Heisenberg
$`\frac{\widehat{L}}{t}=\frac{\widehat{s}}{t}\widehat{\mathrm{\Lambda }}\widehat{s^{}}+\widehat{s}\widehat{\mathrm{\Lambda }}\frac{\widehat{s^{}}}{t}=\frac{i}{\mathrm{}}\widehat{H}\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}}\widehat{\mathrm{\Lambda }}\widehat{s^{}}\frac{i}{\mathrm{}}\widehat{s}\widehat{\lambda }\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}}\widehat{H}`$
$`=\frac{i}{\mathrm{}}(\widehat{H}\widehat{s}\widehat{\mathrm{\Lambda }}\widehat{s^{}}\widehat{s}\widehat{\mathrm{\Lambda }}\widehat{s^{}}\widehat{H})=\frac{i}{\mathrm{}}(\widehat{H}\widehat{L}\widehat{L}\widehat{H})=\frac{i}{\mathrm{}}[\widehat{H},\widehat{L}]`$.
Prin urmare, avem:
$`\frac{\widehat{L}}{t}=\frac{i}{\mathrm{}}[\widehat{H},\widehat{L}]`$.
Deasemenea, ecuaลฃia Heisenberg se poate scrie รฎn forma urmฤtoare:
$`\frac{\widehat{L}}{t}=\frac{i}{\mathrm{}}\widehat{s}[\widehat{H},\widehat{\mathrm{\Lambda }}]\widehat{s^{}}`$.
$`\widehat{L}`$ se cunoaลte ca integralฤ de miลcare dacฤ $`\frac{d}{dt}<\psi \widehat{L}\psi >=0`$ ลi se caracterizeazฤ prin urmฤtorii comutatori:
$`[\widehat{H},\widehat{L}]=0`$, $`[\widehat{H},\widehat{\mathrm{\Lambda }}]=0`$.
1. Stฤri staลฃionare
Stฤrile unui sistem descris prin f.p. ale lui $`\widehat{H}`$ se numesc stฤri staลฃionare ale sistemului, รฎn timp ce setul de v.p. corespunzฤtoare se numeลte spectrul de energie (spectrul energetic) al sistemului. รn astfel de cazuri, ecuaลฃia Schrรถdinger este :
$`i\mathrm{}\frac{\psi _n}{t}=E_n\psi _n=\widehat{H}\psi _n`$.
Soluลฃiile sunt de forma: $`\psi _n(x,t)=\mathrm{exp}^{\frac{iE_nt}{\mathrm{}}}\varphi _n(x)`$.
* Probabilitatea este urmฤtoarea:
$`\delta (x)=\psi _n(x,t)^2=\mathrm{exp}^{\frac{iE_nt}{\mathrm{}}}\varphi _n(x)^2`$
$`=\mathrm{exp}^{\frac{iE_nt}{\mathrm{}}}\mathrm{exp}^{\frac{iE_nt}{\mathrm{}}}\varphi _n(x)^2=\varphi _n(x)^2`$.
Rezultฤ cฤ probabilitatea este constantฤ รฎn timp.
* รn stฤrile staลฃionare, valoarea medie a oricฤrui comutator de forma $`[\widehat{H},\widehat{A}]`$ este zero, unde $`\widehat{A}`$ este un operator arbitrar:
$`<n\widehat{H}\widehat{A}\widehat{A}\widehat{H}n>=<n\widehat{H}\widehat{A}n><n\widehat{A}\widehat{H}n>`$
$`=<nE_n\widehat{A}n><n\widehat{A}E_nn>`$
$`=E_n<n\widehat{A}n>E_n<n\widehat{A}n>=0`$.
* Teorema de virial (virialului) รฎn mecanica cuanticฤ - dacฤ $`\widehat{H}`$ este un operator hamiltonian al unei particule รฎn cรฎmpul $`U(r)`$, folosind
$`\widehat{A}=1/2_{i=1}^3(\widehat{p_i}\widehat{x_i}\widehat{x_i}\widehat{p_i})`$ se obลฃine:
$`<\psi [\widehat{A},\widehat{H}]\psi >=0=<\psi \widehat{A}\widehat{H}\widehat{H}\widehat{A}\psi >`$
$`=_{i=1}^3<\psi \widehat{p_i}\widehat{x_i}\widehat{H}\widehat{H}\widehat{p_i}\widehat{x_i}\psi >`$
$`=_{i=1}^3<\psi [\widehat{H},\widehat{x_i}]\widehat{p_i}+\widehat{x_i}[\widehat{H},\widehat{p_i}]\psi >`$.
folosind de mai multe ori comutatorii ลi $`\widehat{p_i}=i\mathrm{}_i`$, $`\widehat{H}=\widehat{T}+U(r)`$, se poate obลฃine:
$`<\psi [\widehat{A},\widehat{H}]\psi >=0`$
$`=i\mathrm{}(2<\psi \widehat{T}\psi ><\psi \stackrel{}{r}U(r)\psi >)`$.
Aceasta este teorema virialului. Dacฤ potenลฃialul este $`U(r)=U_or^n`$, atunci se obลฃine o formฤ a teoremei de virial ca รฎn mecanica clasicฤ cu unica diferenลฃฤ cฤ se referฤ la valori medii
$`\overline{T}=\frac{n}{2}\overline{U}`$.
* Pentru un Hamiltonian $`\widehat{H}=\frac{\mathrm{}^2}{2m}^2+U(r)`$ ลi $`[\stackrel{}{r},H]=\frac{i\mathrm{}}{m}\stackrel{}{p}`$, calculรฎnd elementele de matrice se obลฃine:
$`(E_kE_n)<n\stackrel{}{r}k>=\frac{i\mathrm{}}{m}<n\widehat{p}k>`$.
1. Densitate de curent de probabilitate Schrรถdinger
Urmฤtoarea integralฤ :
$`\psi _n(x)^2๐x=1`$,
este normalizarea unei f.p. din spectrul discret รฎn reprezentarea de coordonatฤ , ลi apare ca o condiลฃie asupra miลcฤrii microscopice รฎntr-o regiune finitฤ .
Pentru f.p. ale spectrului continuu $`\psi _\lambda (x)`$ nu se poate da รฎn mod direct o interpretare probabilisticฤ .
Sฤ presupunem o funcลฃie datฤ $`\varphi `$ $``$ $`^2`$, pe care o scriem ca o combinaลฃie linearฤ de f.p. รฎn continuu:
$`\varphi =a(\lambda )\psi _\lambda (x)๐x.`$
Se spune cฤ $`\varphi `$ corespunde unei miลcฤri infinite.
รn multe cazuri, funcลฃia $`a(\lambda )`$ este diferitฤ de zero numai รฎntr-o vecinฤtate a unui punct $`\lambda =\lambda _o`$. รn acest caz $`\varphi `$ se cunoaลte ca pachet de unde.
Vom calcula acum viteza de schimbare a probabilitฤลฃii de a gฤsi sistemul รฎn segmentul (volumul) $`\mathrm{\Omega }`$ (generalizarea 3D se face trivial ลi se va presupune, chiar dacฤ notaลฃia este 1D).
$`P=_\mathrm{\Omega }\psi (x,t)^2๐x=_\mathrm{\Omega }\psi ^{}(x,t)\psi (x,t)๐x`$.
Derivรฎnd integrala รฎn raport cu timpul gฤsim:
$`\frac{dP}{dt}=_\mathrm{\Omega }(\psi \frac{\psi ^{}}{t}+\psi ^{}\frac{\psi }{t})๐x`$.
Utilizรฎnd ecuaลฃia Schrรถdinger รฎn integrala din partea dreaptฤ se obลฃine:
$`\frac{dP}{dt}=\frac{i}{\mathrm{}}_\mathrm{\Omega }(\psi \widehat{H}\psi ^{}\psi ^{}\widehat{H}\psi )๐x`$.
Folosind identitatea $`f^2gg^2f=div[(f)grad(g)(g)grad(f)]`$ precum ลi ecuaลฃia Schrรถdinger รฎn forma:
$`\widehat{H}\psi =\frac{\mathrm{}^2}{2m}^2\psi +V\psi `$
ลi substituind รฎn integralฤ se obลฃine:
$`\frac{dP}{dt}=\frac{i}{\mathrm{}}_\mathrm{\Omega }[\psi (\frac{\mathrm{}^2}{2m}^2\psi ^{})\psi ^{}(\frac{\mathrm{}^2}{2m}^2\psi )]๐x`$
$`=_\mathrm{\Omega }\frac{i\mathrm{}}{2m}(\psi ^2\psi ^{}\psi ^{}^2\psi )๐x`$
$`=_\mathrm{\Omega }๐iv\frac{i\mathrm{}}{2m}(\psi ^2\psi ^{}\psi ^{}^2\psi )๐x`$.
Folosind teorema divergenลฃei pentru a transforma integrala de volum รฎn una de suprafaลฃฤ obลฃinem:
$`\frac{dP}{dt}=\frac{i\mathrm{}}{2m}(\psi \psi ^{}\psi ^{}\psi )๐x`$.
Mฤrimea $`\stackrel{}{J}(\psi )=\frac{i\mathrm{}}{2m}(\psi \psi ^{}\psi ^{}\psi )`$ se cunoaลte ca densitate de curent de probabilitate, pentru care imediat se obลฃine o ecuaลฃie de continuitate,
$`\frac{d\rho }{dt}+div(\stackrel{}{J})=0`$.
* Dacฤ $`\psi (x)=AR(x)`$, unde R(x) este o funcลฃie realฤ , atunci: $`\stackrel{}{J}(\psi )=0`$.
* Pentru f.p. ale impulsului $`\psi (x)=\frac{1}{(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{x}}{\mathrm{}}}`$ se obลฃine:
$`J(\psi )=\frac{i\mathrm{}}{2m}(\frac{1}{(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{x}}{\mathrm{}}}(\frac{i\stackrel{}{p}}{\mathrm{}(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{x}}{\mathrm{}}})`$
$`(\frac{1}{(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{x}}{\mathrm{}}}\frac{i\stackrel{}{p}}{\mathrm{}(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\mathrm{}\stackrel{}{p}\stackrel{}{x}}{\mathrm{}}}))`$
$`=\frac{i\mathrm{}}{2m}(\frac{2i\stackrel{}{p}}{\mathrm{}(2\pi \mathrm{})^3})=\frac{\stackrel{}{p}}{m(2\pi \mathrm{})^3}`$,
ceea ce ne indicฤ o densitatea de curent de probabilitate independentฤ de coordonatฤ .
1. Operator de transport spaลฃial
Dacฤ $`\widehat{H}`$ este invariant la translaลฃii de vector arbitrar $`\stackrel{}{a}`$,
$`\widehat{H}(\stackrel{}{r}+\stackrel{}{a})=\widehat{H}\stackrel{}{(r)}`$ ,
atunci existฤ un $`\widehat{T}(\stackrel{}{a})`$ unitar $`\widehat{T}^{}(\stackrel{}{a})\widehat{H}(\stackrel{}{r})\widehat{T}(\stackrel{}{a})=\widehat{H}(\stackrel{}{r}+\stackrel{}{a})`$.
Din cauza comutativitฤลฃii translaลฃiilor
$`\widehat{T}(\stackrel{}{a})\widehat{T}(\stackrel{}{b})=\widehat{T}(\stackrel{}{b})\widehat{T}(\stackrel{}{a})=\widehat{T}(\stackrel{}{a}+\stackrel{}{b})`$,
rezultฤ cฤ $`\widehat{T}`$ are forma $`\widehat{T}=\mathrm{exp}^{i\widehat{k}a}`$, unde $`\widehat{k}=\frac{\widehat{p}}{\mathrm{}}`$.
รn cazul infinitezimal:
$`\widehat{T}(\delta \stackrel{}{a})\widehat{H}\widehat{T}(\delta \stackrel{}{a})(\widehat{I}+i\widehat{k}\delta \stackrel{}{a})\widehat{H}(\widehat{I}i\widehat{k}\delta \stackrel{}{a})`$,
$`\widehat{H}(\stackrel{}{r})+i[\widehat{K},\widehat{H}]\delta \stackrel{}{a}=\widehat{H}(\stackrel{}{r})+(\widehat{H})\delta \stackrel{}{a}`$.
Deasemenea $`[\widehat{p},\widehat{H}]=0`$, unde $`\widehat{p}`$ este o integralฤ de miลcare. Sistemul are funcลฃii de undฤ de forma $`\psi (\stackrel{}{p},\stackrel{}{r})=\frac{1}{(2\pi \mathrm{})^3/2}\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{r}}{\mathrm{}}}`$ ลi transformarea unitarฤ face ca $`\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{a}}{\mathrm{}}}\psi (\stackrel{}{r})=\psi (\stackrel{}{r}+\stackrel{}{a})`$. Operatorul de transport spaลฃial $`\widehat{T}^{}=\mathrm{exp}^{\frac{i\stackrel{}{p}\stackrel{}{a}}{\mathrm{}}}`$ este analogul lui $`\widehat{s}^{}=\mathrm{exp}^{\frac{i\widehat{H}t}{\mathrm{}}}`$, operatorul de โtransportโ temporal.
1. Exemplu: Hamiltonian cristalin
Dacฤ $`\widehat{H}`$ este invariant pentru o translaลฃie discretฤ (de exemplu รฎntr-o reลฃea cristalinฤ ) $`\widehat{H}(\stackrel{}{r}+\stackrel{}{a})=\widehat{H}(\stackrel{}{r})`$, unde $`\stackrel{}{a}=_i\stackrel{}{a_i}n_i`$, $`n_i`$ $``$ $`N`$ ลi $`a_i`$ sunt vectorii barici, atunci:
$`\widehat{H}(\stackrel{}{r})\psi (\stackrel{}{r})=E\psi (\stackrel{}{r})`$,
$`\widehat{H}(\stackrel{}{r}+\stackrel{}{a})\psi (\stackrel{}{r}+\stackrel{}{a})=E\psi (\stackrel{}{r}+\stackrel{}{a})=\widehat{H}(\stackrel{}{r})\psi (\stackrel{}{r}+\stackrel{}{a})`$.
Rezultฤ cฤ $`\psi (\stackrel{}{r})`$ ลi $`\psi (\stackrel{}{r}+\stackrel{}{a})`$ sunt funcลฃii de undฤ pentru aceeaลi v.p. a lui $`\widehat{H}`$. Relaลฃia รฎntre $`\psi (\stackrel{}{r})`$ ลi $`\psi (\stackrel{}{r}+\stackrel{}{a})`$ se poate cฤuta รฎn forma $`\psi (\stackrel{}{r}+\stackrel{}{a})=\widehat{c}(\stackrel{}{a})\psi (\stackrel{}{r})`$ unde $`\widehat{c}(\stackrel{}{a})`$ este o matrice gxg (g este gradul de degenerare al nivelului E). Douฤ matrici de tip coloanฤ , $`\widehat{c}(\stackrel{}{a})`$ $`\widehat{c}(\stackrel{}{b})`$ comutฤ ลi atunci sunt diagonalizabili simultan.
รn plus, pentru elementele diagonale se respectฤ $`c_{ii}(\stackrel{}{a})c_{ii}(\stackrel{}{b})=c_{ii}(\stackrel{}{a}+\stackrel{}{b})`$, i=1,2,โฆ.,g, cu soluลฃii de tipul $`c_{ii}(a)=\mathrm{exp}^{ik_ia}`$. Rezultฤ cฤ $`\psi _k(\stackrel{}{r})=U_k(\stackrel{}{r})\mathrm{exp}^{i\stackrel{}{k}\stackrel{}{a}}`$, unde $`\stackrel{}{k}`$ este un vector real arbitrar ลi funcลฃia $`U_k(\stackrel{}{r})`$ este periodicฤ de perioadฤ $`\stackrel{}{a}`$, $`U_k(\stackrel{}{r}+\stackrel{}{a})=U_k(\stackrel{}{r})`$.
Afirmaลฃia cฤ funcลฃiile proprii ale unui $`\widehat{H}`$ periodic cristalin $`\widehat{H}(\stackrel{}{r}+\stackrel{}{a})=\widehat{H}(\stackrel{}{r})`$ se pot scrie $`\psi _k(\stackrel{}{r})=U_k(\stackrel{}{r})\mathrm{exp}i\stackrel{}{k}\stackrel{}{a}`$ cu $`U_k(\stackrel{}{r}+\stackrel{}{a})=U_k(\stackrel{}{r})`$ se cunoaลte ca teorema lui Bloch. รn cazul continuu, $`U_k`$ trebuie sฤ fie constant, pentru cฤ o constantฤ este unica funcลฃie periodicฤ pentru orice $`\stackrel{}{a}`$. Vectorul $`\stackrel{}{p}=\mathrm{}\stackrel{}{k}`$ se numeลte cuasi-impuls (prin analogie cu cazul continuu). Vectorul $`\stackrel{}{k}`$ nu este determinat รฎn mod univoc, pentru cฤ i se poate adฤuga orice vector $`\stackrel{}{g}`$ pentru care $`ga=2\pi n`$ unde n $``$ N.
Vectorul $`\stackrel{}{g}`$ se poate scrie $`\stackrel{}{g}=_{i=1}^3\stackrel{}{b_i}m_i`$ unde $`m_i`$ sunt numere รฎntregi ลi $`b_i`$ sunt daลฃi de
$`\stackrel{}{b_i}=2\pi \frac{\widehat{a_j}\times \stackrel{}{a_k}}{\stackrel{}{a_i}(\stackrel{}{a_j}\times \stackrel{}{a_k})}`$
pentru $`ijk`$. $`\stackrel{}{b_i}`$ sunt vectorii barici ai reลฃelei cristaline.
Referinลฃe recomandate
1. E. Farhi, J. Goldstone, S. Gutmann, โHow probability arises in quantum mechanicsโ, Annals of Physics 192, 368-382 (1989)
2. N.K. Tyagi รฎn Am. J. Phys. 31, 624 (1963) dฤ o demostraลฃie foarte scurtฤ pentru principiul de incertitudine Heisenberg, pe baza cฤruia se afirmฤ cฤ mฤsurarea simultanฤ a doi operatori hermitici care nu comutฤ produce o incertitudine relaลฃionatฤ cu valoarea comutatorului lor.
3. H.N. Nรบรฑez-Yรฉpez et al., โSimple quantum systems in the momentum representationโ, physics/0001030 (Europ. J. Phys., 2000).
4. J.C. Garrison, โQuantum mechanics of periodic systemsโ, Am. J. Phys. 67, 196 (1999).
5. F. Gieres, โDiracโs formalism and mathematical surprises in quantum mechanicsโ, quant-ph/9907069 (รฎn englezฤ); quant-ph/9907070 (รฎn francezฤ).
1N. Note
1. Pentru โcrearea mecanicii cuanticeโฆโ, Werner Heisenberg a fost distins cu premiul Nobel รฎn 1932 (primit รฎn 1933). Articolul โZรผr Quantenmechanik. IIโ, \[โAsupra mecanicii cuantice.IIโ, Zf. f. Physik 35, 557-615 (1926) (ajuns la redacลฃie รฎn ziua de 16 Noiembrie 1925) de M. Born, W. Heisenberg ลi P. Jordan, se cunoaลte ca โlucrarea celor trei (oameni)โ ลi este considerat ca cel care a deschis cu adevฤrat vastele orizonturi ale mecanicii cuantice.
2. Pentru โinterpretarea statisticฤ a funcลฃiei de undฤ โ Max Born a primit premiul Nobel รฎn 1954.
## 1P. Probleme
Problema 1.1: Se considerฤ doi operatori A ลi B care prin ipotezฤ comutฤ. รn acest caz se poate deduce relaลฃia:
$`e^Ae^B=e^{(A+B)}e^{(1/2[A,B])}`$.
Soluลฃie
Definim un operator F(t), ca funcลฃie de variabilฤ realฤ t, prin: $`F(t)=e^{(At)}e^{(Bt)}`$.
Atunci: $`\frac{dF}{dt}=Ae^{At}e^{Bt}+e^{At}Be^{Bt}=(A+e^{At}Be^{At})F(t)`$.
Acum, aplicรฎnd formula $`[A,F(B)]=[A,B]F^{^{}}(B)`$, avem
$`[e^{At},B]=t[A.B]e^{At}`$, ลi deci: $`e^{At}B=Be^{At}+t[A,B]e^{At},`$
multiplicรฎnd ambele pฤrลฃi ale ecuaลฃiei ultime cu $`\mathrm{exp}^{At}`$ ลi substituind รฎn prima ecuaลฃie, obลฃinem:
$`\frac{dF}{dt}=(A+B+t[A,B])F(t)`$.
Operatorii A, B ลi \[A,B\] comutฤ prin ipotezฤ . Deci, putem integra ecuaลฃia diferenลฃialฤ ca ลi cum $`A+B`$ ลi $`[A,B]`$ ar fi numere (scalare).
Vom avea deci:
$`F(t)=F(0)e^{(A+B)t+1/2[A,B]t^2}`$.
Punรฎnd $`t=0`$, se vede cฤ $`F(0)=1`$, ลi :
$`F(t)=e^{(A+B)t+1/2[A,B]t^2}`$.
Punรฎnd acum $`t=1`$, obลฃinem rezultatul dorit.
Problema 1.2: Sฤ se calculeze comutatorul $`[X,D_x]`$.
Soluลฃie
Calculul se face aplicรฎnd comutatorul unei funcลฃii arbitrare $`\psi (\stackrel{}{r})`$:
$`[X,D_x]\psi (\stackrel{}{r})=(x\frac{}{x}\frac{}{x}x)\psi (\stackrel{}{r})=x\frac{}{x}\psi (\stackrel{}{r})\frac{}{x}[x\psi (\stackrel{}{r})]=x\frac{}{x}\psi (\stackrel{}{r})\psi (\stackrel{}{r})x\frac{}{x}\psi (\stackrel{}{r})=\psi (\stackrel{}{r})`$.
Cum aceastฤ relaลฃie se satisface pentru orice $`\psi (\stackrel{}{r})`$, se poate deduce cฤ : $`[X,D_x]=1`$.
Problema 1.3: Sฤ se verifice cฤ urma de matrice este invariantฤ la schimbฤri de baze ortonormale discrete.
Soluลฃie
Suma elementelor diagonale ale unei reprezentฤri matriceale de un operator (cuantic) A รฎntr-o bazฤ arbitrarฤ nu depinde de bazฤ .
Se va obลฃine aceastฤ propietate pentru cazul schimbฤrii dintr-o bazฤ ortonormalฤ dicretฤ $`u_i>`$ รฎn alta ortonormal discretฤ $`t_k>`$. Avem:
$`_i<u_iAu_i>=_i<u_i\left(_kt_k><t_k\right)Au_i>`$
(unde s-a folosit relaลฃia de completitudine pentru starea $`t_k`$). Partea dreaptฤ este egalฤ cu:
$`_{i,j}<u_it_k><t_kAu_i>=_{i,j}<t_kAu_i><u_it_k>`$,
(este posibilฤ schimbarea ordinii รฎn produsul de douฤ numere scalare). Astfel, putem รฎnlocui $`_iu_i><u_i`$ cu unu (relaลฃia de completitudine pentru stฤrile $`u_i>`$), pentru a obลฃine รฎn final:
$$\underset{i}{}<u_iAu_i>=\underset{k}{}<t_kAt_k>.$$
Aลadar, s-a demonstrat proprietatea cerutฤ de invarianลฃฤ pentru urmele matriceale.
Problema 1.4: Dacฤ pentru operatorul hermitic $`N`$ existฤ operatorii hermitici $`L`$ ลi $`M`$ astfel cฤ : $`[M,N]=0`$, $`[L,N]=0`$, $`[M,L]0`$, atunci funcลฃiile proprii ale lui $`N`$ sunt degenerate.
Soluลฃie
Fie $`\psi (x;\mu ,\nu )`$ funcลฃiile proprii comune ale lui $`M`$ ลi $`N`$ (fiind de comutator nul sunt observabile simultane). Fie $`\psi (x;\lambda ,\nu )`$ funcลฃiile proprii comune ale lui $`L`$ ลi $`N`$ (fiind de comutator nul sunt observabile simultane). Parametrii greci indicฤ valorile proprii ale operatorilor corespunzฤtori. Considerฤm, pentru a simplifica, cฤ $`N`$ are spectru discret. Atunci:
$$f(x)=\underset{\nu }{}a_\nu \psi (x;\mu ,\nu )=\underset{\nu }{}b_\nu \psi (x;\lambda ,\nu ).$$
Se calculeazฤ acum elementul de matrice $`<f|ML|f>`$:
$$<f|ML|f>=\underset{\nu }{}\mu _\nu a_\nu \psi ^{}(x;\mu ,\nu )\underset{\nu ^{^{}}}{}\lambda _\nu ^{^{}}b_\nu ^{^{}}\psi (x;\lambda ,\nu ^{^{}})dx.$$
Dacฤ toate funcลฃiile proprii ale lui $`N`$ sunt diferite (nedegenerate) atunci $`<f|ML|f>=_\nu \mu _\nu a_\nu \lambda _\nu b_\nu `$. Dar acelaลi rezultat se obลฃine ลi dacฤ se calculeazฤ $`<f|LM|f>`$ ลi comuatorul ar fi zero. Prin urmare, unele funcลฃii proprii ale lui $`N`$ trebuie sฤ fie degenerate.
## 2. BARIERE ลI GROPI RECTANGULARE
### Regiuni de potenลฃial constant
รn cazul unui potenลฃial rectangular, $`V(x)`$ este o funcลฃie constantฤ $`V(x)=V`$ รฎntr-o regiune oarecare รฎn spaลฃiu. รntr-o astfel de regiune, ecuaลฃia Schrรถdinger poate fi scrisฤ :
$$\frac{d^2\psi (x)}{dx^2}+\frac{2m}{\mathrm{}^2}(EV)\psi (x)=0$$
(1)
Distingem mai multe cazuri:
(i) $`E>V`$
Introducem constanta pozitivฤ $`k`$, definitฤ prin
$$k=\frac{\sqrt{2m(EV)}}{\mathrm{}}$$
(2)
Soluลฃia ecuaลฃiei (1) se poate atunci scrie:
$$\psi (x)=Ae^{ikx}+A^{}e^{ikx}$$
(3)
unde $`A`$ ลi $`A^{}`$ sunt constante complexe.
(ii) $`E<V`$
Aceastฤ condiลฃie corespunde la regiuni de spaลฃiu care ar fi interzise pentru particulฤ din punctul de vedere al miลcฤrii mecanice clasice. รn acest caz, introducem constanta pozitivฤ $`q`$ definitฤ prin:
$$q=\frac{\sqrt{2m(VE)}}{\mathrm{}}$$
(4)
ลi soluลฃia lui (1) poate fi scrisฤ :
$$\psi (x)=Be^{qx}+B^{}e^{qx}$$
(5)
unde $`B`$ ลi $`B^{}`$ sunt constante complexe.
(iii) $`E=V`$
รn acest caz special, $`\psi (x)`$ este o funcลฃie linearฤ de $`x`$.
### Comportamentul lui $`\psi (x)`$ la o discontinuitate a potenลฃialului
S-ar putea crede cฤ in punctul $`x=x_1`$, unde potenลฃialul $`V(x)`$ este discontinuu, funcลฃia de undฤ $`\psi (x)`$ se comportฤ mai ciudat, poate cฤ รฎn mod discontinuu, de exemplu. Aceasta nu se รฎntรฎmplฤ : $`\psi (x)`$ ลi $`\frac{d\psi }{dx}`$ sunt continue, ลi numai a doua derivatฤ prezintฤ discontinuitate รฎn $`x=x_1`$.
### Viziune generalฤ asupra calculului
Procedeul pentru determinarea stฤrii staลฃionare รฎn potenลฃiale rectangulare este deci urmฤtorul: รฎn toate regiunile unde $`V(x)`$ este constant, scriem $`\psi (x)`$ รฎn oricare dintre cele douฤ forme (3) sau (5) รฎn funcลฃie de aplicaลฃie; รฎn continuare โlipimโ aceste funลฃii corespunzฤtor cerinลฃei de continuitate pentru $`\psi (x)`$ ลi $`\frac{d\psi }{dx}`$ รฎn punctele unde $`V(x)`$ este discontinuu.
## Examinarea cรฎtorva cazuri simple
Sฤ facem calculul cantitativ pentru stฤrile staลฃionare, conform metodei descrise.
### Potenลฃial treaptฤ
a. Cazul $`E>V_0`$; reflexie parลฃialฤ
Sฤ punem ec. (2) รฎn forma:
$`k_1`$ $`=`$ $`{\displaystyle \frac{\sqrt{2mE}}{\mathrm{}}}`$ (6)
$`k_2`$ $`=`$ $`{\displaystyle \frac{\sqrt{2m(EV_0)}}{\mathrm{}}}`$ (7)
Soluลฃia ec. (1) are forma din ec. (3) รฎn regiunile $`I(x<0)`$ ลi $`II(x>0)`$:
$`\psi _I`$ $`=`$ $`A_1e^{ik_1x}+A_1^{}e^{ik_1x}`$
$`\psi _{II}`$ $`=`$ $`A_2e^{ik_2x}+A_2^{}e^{ik_2x}`$
รn regiunea I ec. (1) ia forma:
$`\psi ^{\prime \prime }(x)+{\displaystyle \frac{2mE}{\mathrm{}^2}}\psi (x)=\psi ^{\prime \prime }(x)+k^2\psi (x)=0`$
iar รฎn regiunea II:
$`\psi ^{\prime \prime }(x){\displaystyle \frac{2m}{\mathrm{}^2}}[V_0E]\varphi (x)=\psi ^{\prime \prime }(x)q^2\psi (x)=0`$
Dacฤ ne limitฤm la cazul unei particule incidente care โvineโ de la $`x=\mathrm{}`$, trebuie sฤ alegem $`A_2^{}=0`$ ลi putem determina raporturile $`A_1^{}/A_1`$ ลi $`A_2/A_1`$. Condiลฃiile de โlipireโ dau atunci:
* $`\psi _I=\psi _{II}`$, รฎn $`x=0:`$
$$A_1+A_1^{}=A_2$$
(8)
* $`\psi _I^{}=\psi _{II}^{}`$, รฎn $`x=0:`$
$$A_1ik_1A_1^{}ik_1=A_2ik_2$$
(9)
Substituind $`A_1`$ ลi $`A_1^{}`$ din (8) รฎn (9):
$`A_1^{}`$ $`=`$ $`{\displaystyle \frac{A_2(k_1k_2)}{2k_1}}`$ (10)
$`A_1`$ $`=`$ $`{\displaystyle \frac{A_2(k_1+k_2)}{2k_1}}`$ (11)
Egalarea constantei $`A_2`$ รฎn (10) ลi (11) implicฤ :
$$\frac{A_1^{}}{A_1}=\frac{k_1k_2}{k_1+k_2}$$
(12)
ลi din (11) obลฃinem:
$$\frac{A_2}{A_1}=\frac{2k_1}{k_1+k_2}$$
(13)
$`\psi (x)`$ este o superpoziลฃia de douฤ unde. Prima (termenul รฎn $`A_1`$) corespunde unei particule incidente, de moment $`p=\mathrm{}k_1`$, รฎn propagare de la stรฎnga la dreapta. A doua (termenul รฎn $`A_1^{}`$) corespunde unei particule reflectate, de impuls $`\mathrm{}k_1`$, รฎn propagare รฎn sens opus. Cum deja am ales $`A_2^{}=0`$, $`\psi _{II}(x)`$ conลฃine o singurฤ undฤ , care este asociatฤ cu o particulฤ transmisฤ . (Se va arฤta mai departe cum este posibil, folosind conceptul de curent de probabilitate, sฤ definim coeficientul de transmisie T precum ลi coeficientul de reflexie R pentru potenลฃialul treaptฤ ). Aceลti coeficienลฃi dau probabilitatea ca o particulฤ , sosind de la $`x=\mathrm{}`$, ar putea trece de potenลฃialul treaptฤ รฎn $`x=0`$ sau se รฎntoarce. Astfel obลฃinem:
$$R=|\frac{A_1^{}}{A_1}|^2$$
(14)
iar pentru $`T`$:
$$T=\frac{k_2}{k_1}|\frac{A_2}{A_1}|^2.$$
(15)
ลขinรฎnd cont de (12) ลi (13), avem:
$`R`$ $`=`$ $`1{\displaystyle \frac{4k_1k_2}{(k_1+k_2)^2}}`$ (16)
$`T`$ $`=`$ $`{\displaystyle \frac{4k_1k_2}{(k_1+k_2)^2}}.`$ (17)
Este uลor de verificat cฤ $`R+T=1`$: este deci sigur cฤ particula va fi transmisฤ sau reflectatฤ . Contrar predicลฃiilor mecanicii clasice, particula incidentฤ are o probabilitate nenulฤ de a nu se รฎntoarce.
Deasemenea este uลor de verificat, folosind (6), (7) ลi (17), cฤ dacฤ $`EV_0`$ atunci $`T1`$: cรฎnd energia particulei este suficient de mare รฎn comparaลฃie cu รฎnฤlลฃimea treptei, pentru particulฤ este ca ลi cum obstacolul treaptฤ nu ar exista.
Considerรฎnd soluลฃia รฎn regiunea I:
$`\psi _I=A_1e^{ik_1x}+A_1^{}e^{ik_1x}`$
$$j=\frac{i\mathrm{}}{2m}(\varphi ^{}\varphi \varphi \varphi ^{})$$
(18)
cu $`A_1e^{ik_1x}`$ ลi conjugata sa $`A_1^{}e^{ik_1x}`$:
$`j`$ $`=`$ $`{\displaystyle \frac{i\mathrm{}}{2m}}[(A_1^{}e^{ik_1x})(A_1ik_1e^{ik_1x})(A_1e^{ik_1x})(A_1^{}ik_1e^{ik_1x})]`$
$`j`$ $`=`$ $`{\displaystyle \frac{\mathrm{}k_1}{m}}|A_1|^2`$
Acum cu $`A_1^{}e^{ik_1x}`$ ลi conjugata sa $`A_1^{}e^{ik_1x}`$ rezultฤ :
$`j={\displaystyle \frac{\mathrm{}k_1}{m}}|A_1^{}|^2.`$
Dorim รฎn continuare sฤ verificฤm proporลฃia de curent reflectat faลฃฤ de curentul incident (mai precis, dorim sฤ verificฤm probabilitatea ca particula sฤ fie returnatฤ ):
$`R={\displaystyle \frac{|j(\varphi _{})|}{|j(\varphi _+)|}}={\displaystyle \frac{|\frac{\mathrm{}k_1}{m}|A_1^{}|^2|}{|\frac{\mathrm{}k_1}{m}|A_1|^2|}}=|{\displaystyle \frac{A_1^{}}{A_1}}|^2`$ (19)
Similar, proporลฃia de transmisie faลฃฤ de incidenลฃฤ (adicฤ probabilitatea ca particula sฤ fie transmisฤ ) este, ลฃinรฎnd acum cont de soluลฃia din regiunea II:
$`T={\displaystyle \frac{|\frac{\mathrm{}k_2}{m}|A_2|^2|}{|\frac{\mathrm{}k_1}{m}|A_1|^2|}}={\displaystyle \frac{k_2}{k_1}}|{\displaystyle \frac{A_2}{A_1}}|^2`$ (20)
a. Cazul $`E<V_0`$; reflexie totalฤ
รn acest caz avem:
$`k_1`$ $`=`$ $`{\displaystyle \frac{\sqrt{2mE}}{\mathrm{}}}`$ (21)
$`q_2`$ $`=`$ $`{\displaystyle \frac{\sqrt{2m(V_0E)}}{\mathrm{}}}`$ (22)
รn regiunea $`I(x<0)`$, soluลฃia ec. (1) \[scrisฤ $`\psi (x)^{\prime \prime }+k_1^2\psi (x)=0`$\] are forma datฤ รฎn ec. (3):
$$\psi _I=A_1e^{ik_1x}+A_1^{}e^{ik_1x}$$
(23)
iar รฎn regiunea $`II(x>0)`$, aceeaลi ec. (1) \[acum scrisฤ $`\psi (x)^{\prime \prime }q_2^2\psi (x)=0`$\] are forma ec. (5):
$$\psi _{II}=B_2e^{q_2x}+B_2^{}e^{q_2x}$$
(24)
Pentru ca soluลฃia sฤ fie menลฃinutฤ finitฤ cรฎnd $`x+\mathrm{}`$, este necesar ca:
$$B_2=0$$
(25)
Condiลฃiile de โlipitโ รฎn $`x=0`$ dau รฎn acest caz:
* $`\psi _I=\psi _{II}`$, รฎn $`x=0:`$
$$A_1+A_1^{}=B_2^{}$$
(26)
* $`\psi _I^{}=\psi _{II}^{}`$, รฎn $`x=0:`$
$$A_1ik_1A_1^{}ik_1=B_2^{}q_2$$
(27)
Substituind $`A_1`$ ลi $`A_1^{}`$ din (26) รฎn (27):
$`A_1^{}`$ $`=`$ $`{\displaystyle \frac{B_2^{}(ik_1+q_2)}{2ik_1}}`$ (28)
$`A_1`$ $`=`$ $`{\displaystyle \frac{B_2^{}(ik_1q_2)}{2ik1}}`$ (29)
Egalarea constantei $`B_2^{}`$ รฎn (28) ลi (29) duce la:
$$\frac{A_1^{}}{A_1}=\frac{ik_1+q_2}{ik_1q_2}=\frac{k_1iq_2}{k_1+iq_2},$$
(30)
astfel cฤ din (29) avem:
$$\frac{B_2^{}}{A_1}=\frac{2ik_1}{ik_1q_2}=\frac{2k_1}{k_1iq_2}$$
(31)
Coeficientul de reflexie $`R`$ este deci:
$$R=|\frac{A_1^{}}{A_1}|^2=|\frac{k_1iq_2}{k_1+iq_2}|^2=\frac{k_1^2+q_2^2}{k_1^2+q_2^2}=1$$
(32)
Ca รฎn mecanica clasicฤ , microparticula este รฎntotdeauna reflectatฤ (reflexie totalฤ ). Totuลi, existฤ o diferenลฃฤ importantฤ : datoritฤ existenลฃei aลa-numitei unde evanescente $`e^{q_2x}`$, particula are o probabilitate nenulฤ de a se gฤsi โprezentฤ โ รฎn regiunea din spaลฃiu care este clasic interzisฤ . Aceastฤ probabilitate descreลte exponenลฃial cu $`x`$ ลi ajunge sฤ fie neglijabilฤ cรฎnd $`x`$ depฤลeลte โzonaโ $`1/q_2`$ corespunzฤtoare undei evanescente. Sฤ notฤm ลi cฤ $`A_1^{}/A_1`$ este complex. O anumitฤ diferenลฃฤ de fazฤ apare din cauza reflexiei, care, fizic, se datoreazฤ faptului cฤ particula este โ frรฎnatฤโ (รฎncetinitฤ cรฎnd intrฤ รฎn regiunea $`x>0`$. Nu existฤ analogie pentru aceasta รฎn mecanica clasicฤ , ci doar รฎn optica fizicฤ .
### Potenลฃiale tip barierฤ
a. Cazul $`E>V_0`$; rezonanลฃe
Sฤ punem aici ec. (2) รฎn forma:
$`k_1`$ $`=`$ $`{\displaystyle \frac{\sqrt{2mE}}{\mathrm{}}}`$ (33)
$`k_2`$ $`=`$ $`{\displaystyle \frac{\sqrt{2m(EV_0)}}{\mathrm{}}}`$ (34)
Soluลฃia ec. (1) este ca รฎn ec. (3) รฎn regiunile $`I(x<0)`$, $`II(0<x<a`$) ลi $`III(x>a):`$
$`\psi _I`$ $`=`$ $`A_1e^{ik_1x}+A_1^{}e^{ik_1x}`$
$`\psi _{II}`$ $`=`$ $`A_2e^{ik_2x}+A_2^{}e^{ik_2x}`$
$`\psi _{III}`$ $`=`$ $`A_3e^{ik_1x}+A_3^{}e^{ik_1x}`$
Dacฤ ne limitฤm la cazul unei particule incidente care vine de la $`x=\mathrm{}`$, trebuie sฤ alegem $`A_3^{}=0`$.
* $`\psi _I=\psi _{II}`$, รฎn $`x=0:`$
$$A_1+A_1^{}=A_2+A_2^{}$$
(35)
* $`\psi _I^{}=\psi _{II}^{}`$, รฎn $`x=0:`$
$$A_1ik_1A_1^{}ik_1=A_2ik_2A_2^{}ik_2$$
(36)
* $`\psi _{II}=\psi _{III}`$, รฎn $`x=a:`$
$$A_2e^{ik_2a}+A_2^{}e^{ik_2a}=A_3e^{ik_1a}$$
(37)
* $`\psi _{II}^{}=\psi _{III}^{}`$, รฎn $`x=a:`$
$$A_2ik_2e^{ik_2a}A_2^{}ik_2e^{ik_2a}=A_3ik_1e^{ik_1a}$$
(38)
Condiลฃiile de continuitate รฎn $`x=a`$ dau $`A_2`$ ลi $`A_2^{}`$ รฎn funcลฃie de $`A_3`$, ลi cele din $`x=0`$ dau $`A_1`$ ลi $`A_1^{}`$ รฎn funลฃie de $`A_2`$ ลi $`A_2^{}`$ (ลi deci รฎn funcลฃie de $`A_3`$). Acest procedeu este arฤtat รฎn continuare.
Substituind $`A_2^{}`$ din ec. (37) รฎn (38):
$$A_2=\frac{A_3e^{ik_1a}(k_2+k_1)}{2k_2e^{ik_2a}}$$
(39)
Substituind $`A_2`$ din ec. (37) รฎn (38):
$$A_2^{}=\frac{A_3e^{ik_1a}(k_2k_1)}{2k_2e^{ik_2a}}$$
(40)
Substituind $`A_1`$ din ec. (35) รฎn (36):
$$A_1^{}=\frac{A_2(k_2k_1)A_2^{}(k_2+k_1)}{2k_1}$$
(41)
Substituind $`A_1^{}`$ din ec. (35) รฎn (36):
$$A_1=\frac{A_2(k_2+k_1)A_2^{}(k_2k_1)}{2k_1}$$
(42)
Acum, substituind รฎn (41) ecuaลฃiile (39) ลi (40), avem:
$$A_1^{}=i\frac{(k_2^2k_1^2)}{2k_1k_2}(\mathrm{sin}k_2a)e^{ik_1a}A_3$$
(43)
รn final, substituind รฎn (42) ecuaลฃiile (39) ลi (40):
$$A_1=[\mathrm{cos}k_2ai\frac{k_1^2+k_2^2}{2k_1k_2}\mathrm{sin}k_2a]e^{ik_1a}A_3$$
(44)
$`A_1^{}/A_1`$ ลi $`A_3/A_1`$ \[raporturi care se obลฃin egalรฎnd ecuaลฃiile (43) ลi (44), ลi respectiv separรฎnd รฎn ec. (44)\] ne permit calculul coeficientului de reflexie $`R`$ precum ลi a celui de transmisie $`T`$ pentru acest caz simplu de barierฤ , fiind aceลtia daลฃi de:
$$R=|A_1^{}/A_1|^2=\frac{(k_1^2k_2^2)^2\mathrm{sin}^2k_2a}{4k_1^2k_2^2+(k_1^2k_2^2)^2\mathrm{sin}^2k_2a},$$
(45)
$$T=|A_3/A_1|^2=\frac{4k_1^2k_2^2}{4k_1^2k_2^2+(k_1^2k_2^2)^2\mathrm{sin}^2k_2a},$$
(46)
Acum este uลor de verificar cฤ $`R+T=1`$.
b. Cazul $`E<V_0`$; efectul tunel
Acum, fie ecuaลฃiile (2) ลi (4):
$`k_1`$ $`=`$ $`{\displaystyle \frac{\sqrt{2mE}}{\mathrm{}}}`$ (47)
$`q_2`$ $`=`$ $`{\displaystyle \frac{\sqrt{2m(V_0E)}}{\mathrm{}}}`$ (48)
Soluลฃia ec. (1) are forma ec. (3) รฎn regiunile $`I(x<0)`$ ลi $`III(x>a)`$, รฎn timp ce รฎn regiunea $`II(0<x<a`$) are forma ec. (5):
$`\psi _I`$ $`=`$ $`A_1e^{ik_1x}+A_1^{}e^{ik_1x}`$
$`\psi _{II}`$ $`=`$ $`B_2e^{q_2x}+B_2^{}e^{q_2x}`$
$`\psi _{III}`$ $`=`$ $`A_3e^{ik_1x}+A_3^{}e^{ik_1x}`$
Condiลฃiile de โlipitโ รฎn $`x=0`$ ลi $`x=a`$ ne permit calculul coeficientului de transmisie al barierei. De fapt, nu este necesar a efectua รฎncฤ odata calculul: este suficient de a face substituลฃia, รฎn ecuaลฃia obลฃinutฤ รฎn primul caz din aceastฤ secลฃiune $`k_2`$ cu $`iq_2`$.
### Stฤri legate รฎn groapฤ rectangularฤ
a. Groapฤ de adรฎncime finitฤ
รn aceastฤ parte ne limitฤm la studiul cazului $`0<E<V_0`$ (cazul $`E>V_0`$ este identic calculului din secลฃiunea precedentฤ , โbarierฤ de potenลฃialโ.
Pentru regiunile exterioare I $`(x<0)`$ ลi III $`(x>a)`$ folosim ec. (4):
$$q=\frac{\sqrt{2m(V_0E)}}{\mathrm{}}$$
(49)
Pentru regiunea centralฤ II $`(0<x<a)`$ folosim ec. (2):
$$k=\frac{\sqrt{2m(E)}}{\mathrm{}}$$
(50)
Soluลฃia ec. (1) are forma ec. (5) รฎn regiunile exterioare ลi รฎn forma din ec. (3) รฎn regiunea centralฤ :
$`\psi _I`$ $`=`$ $`B_1e^{qx}+B_1^{}e^{qx}`$
$`\psi _{II}`$ $`=`$ $`A_2e^{ikx}+A_2^{}e^{ikx}`$
$`\psi _{III}`$ $`=`$ $`B_3e^{qx}+B_3^{}e^{qx}`$
รn regiunea $`(0<x<a)`$ ec. (1) are forma:
$$\psi (x)^{\prime \prime }+\frac{2mE}{\mathrm{}^2}\psi (x)=\psi (x)^{\prime \prime }+k^2\psi (x)=0$$
(51)
ลi รฎn regiunile exterioare:
$$\psi (x)^{\prime \prime }\frac{2m}{\mathrm{}^2}[V_0E]\varphi (x)=\psi (x)^{\prime \prime }q^2\psi (x)=0$$
(52)
Pentru cฤ $`\psi `$ trebuie sฤ fie finitฤ รฎn regiunea I, trebuie sฤ avem:
$$B_1^{}=0$$
(53)
Condiลฃiile de lipire dau:
$`\psi _I=\psi _{II}`$, รฎn $`x=0:`$
$$B_1=A_2+A_2^{}$$
(54)
$`\psi _I^{}=\psi _{II}^{}`$, รฎn $`x=0:`$
$$B_1q=A_2ikA_2^{}ik$$
(55)
$`\psi _{II}=\psi _{III}`$, รฎn $`x=a:`$
$$A_2e^{ika}+A_2^{}e^{ika}=B_3e^{qa}+B_3^{}e^{qa}$$
(56)
$`\psi _{II}^{}=\psi _{III}^{}`$, รฎn $`x=a:`$
$$A_2ike^{ika}A_2^{}ike^{ika}=B_3qe^{qa}B_3^{}qe^{qa}$$
(57)
Substituind constantele $`A_2`$ ลi $`A_2^{}`$ din ec. (54) รฎn ec. (55) obลฃinem, respectiv:
$`A_2^{}`$ $`=`$ $`{\displaystyle \frac{B_1(qik)}{2ik}}`$
$`A_2`$ $`=`$ $`{\displaystyle \frac{B_1(q+ik)}{2ik}}`$ (58)
Substituind constanta $`A_2`$ ลi constanta $`A_2^{}`$ din ec. (56) รฎn ec. (57) obลฃinem, respectiv:
$`B_3^{}e^{qa}(ik+q)+B_3e^{qa}(ikq)+A_2^{}e^{ika}(2ik)`$ $`=`$ $`0`$
$`2ikA_2e^{ika}+B_3^{}e^{qa}(ik+q)+B_3E^{qa}(ikq)`$ $`=`$ $`0`$ (59)
Egalรฎnd $`B_3^{}`$ din ecuaลฃiile (59) ลi ลฃinรฎnd cont de ecuaลฃiile (58):
$$\frac{B_3}{B_1}=\frac{e^{qa}}{4ikq}[e^{ika}(q+ik)^2e^{ika}(qik)^2]$$
(60)
รnsฤ $`\psi (x)`$ trebuie sฤ fie finitฤ ลi รฎn regiunea III. Prin urmare, este necesar ca $`B_3=0`$, ลi deci:
$$[\frac{qik}{q+ik}]^2=\frac{e^{ika}}{e^{ika}}=e^{2ika}$$
(61)
Deoarece $`q`$ ลi $`k`$ depind de $`E`$, ec. (1) poate fi satisfฤ cutฤ pentru anumite valori ale lui $`E`$. Condiลฃia ca $`\psi (x)`$ sฤ fie finitฤ รฎn toate regiunile spaลฃiale impune cuantizarea energiei. ลi mai precis douฤ cazuri sunt posibile:
(i) dacฤ :
$$\frac{qik}{q+ik}=e^{ika}$$
(62)
Egalรฎnd รฎn ambii membri partea realฤ ลi cea imaginarฤ ,respectiv, rezultฤ :
$$\mathrm{tan}(\frac{ka}{2})=\frac{q}{k}$$
(63)
Punรฎnd:
$$k_0=\sqrt{\frac{2mV_0}{\mathrm{}}}=\sqrt{k^2+q^2}$$
(64)
obลฃinem:
$$\frac{1}{\mathrm{cos}^2(\frac{ka}{2})}=1+\mathrm{tan}^2(\frac{ka}{2})=\frac{k^2+q^2}{k^2}=(\frac{k_0}{k})^2$$
(65)
Ec.(63) este astfel echivalentฤ cu sistemul de ecuaลฃii:
$`|\mathrm{cos}({\displaystyle \frac{ka}{2}})|`$ $`=`$ $`{\displaystyle \frac{k}{k_0}}`$
$`\mathrm{tan}({\displaystyle \frac{ka}{2}})`$ $`>`$ $`0`$ (66)
Nivelele de energie sunt determinate de cฤtre intersecลฃia unei linii drepte de รฎnclinare $`\frac{1}{k_0}`$ cu primul set de cosinusoide รฎntrerupte รฎn figura 2.4. Astfel obลฃinem un numฤr de nivele de energie, ale cฤror funcลฃii de undฤ sunt pare. Acest lucru devine mai clar dacฤ substituim (62) รฎn (58) ลi (60). Este uลor de verificat cฤ $`B_3^{}=B_1`$ ลi $`A_2=A_2^{}`$, astfel cฤ $`\psi (x)=\psi (x)`$.
(ii) dacฤ :
$$\frac{qik}{q+ik}=e^{ika}$$
(67)
Un calcul de acelaลi tip ne duce la:
$`|\mathrm{sin}({\displaystyle \frac{ka}{2}})|`$ $`=`$ $`{\displaystyle \frac{k}{k_0}}`$
$`\mathrm{tan}({\displaystyle \frac{ka}{2}})`$ $`<`$ $`0`$ (68)
Nivelele de energie sunt รฎn acest caz determinate de cฤtre intersecลฃia aceleiaลi linii drepte cu al doilea set de cosinusoide รฎntrerupte รฎn figura 2.4. Nivelele astfel obลฃinute se aflฤ รฎntre cele gฤsite รฎn (i). Se poate arฤta uลor cฤ funcลฃiile de undฤ corespunzฤtoare sunt impare.
b. Groapฤ de adรฎncime infinitฤ
รn acest caz este convenabil sฤ se punฤ $`V(x)`$ zero pentru $`0<x<a`$ ลi infinit รฎn tot restul axei. Punรฎnd:
$$k=\sqrt{\frac{2mE}{\mathrm{}^2}}$$
(69)
$`\psi (x)`$ trebuie sฤ fie zero รฎn afara intervalului $`[0,a]`$, ลi continuฤ รฎn $`x=0`$, cรฎt ลi รฎn $`x=a`$.
Acum, pentru $`0xa`$:
$$\psi (x)=Ae^{ikx}+A^{}e^{ikx}$$
(70)
Pentru cฤ $`\psi (0)=0`$, se poate deduce cฤ $`A^{}=A`$, ceea ce ne conduce la:
$$\psi (x)=2iA\mathrm{sin}(kx)$$
(71)
รn plus $`\psi (a)=0`$, astfel cฤ :
$$k=\frac{n\pi }{a}$$
(72)
unde $`n`$ este un รฎntreg pozitiv arbitrar. Dacฤ normalizฤm funcลฃia (71), ลฃinรฎnd cont de (72), atunci obลฃinem funcลฃiile de undฤ staลฃionare:
$$\psi _n(x)=\sqrt{\frac{2}{a}}\mathrm{sin}(\frac{n\pi x}{a})$$
(73)
cu energiile:
$$E_n=\frac{n^2\pi ^2\mathrm{}^2}{2ma^2}$$
(74)
Cuantizarea nivelelor de energie este prin urmare extrem de simplฤ รฎn acest caz: energiile staลฃionare sunt proporลฃionale cu pฤtratele numerelor naturale.
## 2P. Probleme
### Problema 2.1: Potenลฃialul Delta atractiv
Sฤ presupunem cฤ avem un potenลฃial de forma:
$`V(x)=V_0\delta (x);V_0>0;x\mathrm{}.`$
Funcลฃia de undฤ $`\psi (x)`$ corespunzฤtoare se presupune continuฤ .
a) Sฤ se obลฃinฤ stฤrile legate ($`E<0`$), dacฤ existฤ , localizate รฎn acest tip de potenลฃial.
b) Sฤ se calculeze dispersia unei unde plane care โcadeโ pe acest potenลฃial ลi sฤ se obลฃinฤ coeficientul de reflexie
$`R={\displaystyle \frac{|\psi _{refl}|^2}{|\psi _{inc}|^2}}|_{x=0}`$
unde $`\psi _{refl}`$, $`\psi _{inc}`$ sunt unda reflectatฤ ลi respectiv cea incidentฤ .
Sugestie: Pentru a evalua comportamentul lui $`\psi (x)`$ รฎn x=0, se recomandฤ integrarea ecuaลฃiei Schrรถdinger รฎn intervalul ($`\epsilon ,+\epsilon `$), dupฤ care se aplicฤ limita $`\epsilon `$ $``$ $`0`$.
Soluลฃie. a) Ecuaลฃia Schrรถdinger este:
$$\frac{d^2\psi (x)}{dx^2}+\frac{2m}{\mathrm{}^2}(E+V_0\delta (x))\psi (x)=0$$
(75)
Departe de origine avem o ecuaลฃie diferenลฃialฤ de forma
$$\frac{d^2}{dx^2}\psi (x)=\frac{2mE}{\mathrm{}^2}\psi (x).$$
(76)
Funcลฃiile de undฤ sunt prin urmare de forma
$$\psi (x)=Ae^{qx}+Be^{qx}\mathrm{pentru}x>0\mathrm{sau}x<0,$$
(77)
cu $`q=\sqrt{2mE/\mathrm{}^2}`$ $`\mathrm{}.`$ Cum $`|\psi |^2`$ trebuie sฤ fie integrabilฤ , nu putem accepta o parte care sฤ creascฤ exponenลฃial. รn plus funcลฃia de undฤ trebuie sฤ fie continuฤ รฎn origine. Cu aceste condiลฃii,
$`\psi (x)`$ $`=`$ $`Ae^{qx};(x<0),`$
$`\psi (x)`$ $`=`$ $`Ae^{qx};(x>0).`$ (78)
รntegrรฎnd ecuaลฃia Schrรถdinger รฎntre $`\epsilon `$ ลi $`+\epsilon `$, obลฃinem
$$\frac{\mathrm{}^2}{2m}[\psi ^{}(\epsilon )\psi ^{}(\epsilon )]V_0\psi (0)=E_\epsilon ^{+\epsilon }\psi (x)๐x2\epsilon E\psi (0)$$
(79)
Introducรฎnd acum resultatul (78) ลi ลฃinรฎnd cont de limita $`\epsilon 0`$, avem
$$\frac{\mathrm{}^2}{2m}(qAqA)V_0A=0$$
(80)
sau $`E=m(V_0^2/2\mathrm{}^2)`$ \[$`\frac{V_0^2}{4}`$ รฎn unitฤลฃi $`\frac{\mathrm{}^2}{2m}`$\]. รn mod clar existฤ o singurฤ energie discretฤ . Constanta de normalizare se gฤseลte cฤ este $`A=\sqrt{mV_0/\mathrm{}^2}`$. Funcลฃia de undฤ a stฤrii legate se obลฃine $`\psi _o=Ae^{V_0|x|/2}`$, cu $`V_0`$ รฎn unitฤลฃi $`\frac{\mathrm{}^2}{2m}`$. b) Funcลฃia de undฤ pentru o undฤ planฤ este dupฤ cum se ลtie
$$\psi (x)=Ae^{ikx},k^2=\frac{2mE}{\mathrm{}^2}.$$
(81)
Se miลcฤ de la stรฎnga la dreapta ลi se reflectฤ รฎn potenลฃial. Dacฤ $`B`$ sau $`C`$ este amplitudinea undei reflectate sau transmise, respectiv, avem
$`\psi (x)`$ $`=`$ $`Ae^{ikx}+Be^{ikx};(x<0),`$
$`\psi (x)`$ $`=`$ $`Ce^{ikx};(x>0).`$ (82)
Condiลฃiile de continuitate ลi relaลฃia $`\psi ^{}(\epsilon )\psi ^{}(\epsilon )=f\psi (0)`$ cu $`f=2mV_0/\mathrm{}^2`$ produc
$`A+B`$ $`=`$ $`CB={\displaystyle \frac{f}{f+2ik}}A,`$
$`ik(CA+B)`$ $`=`$ $`fCC={\displaystyle \frac{2ik}{f+2ik}}A.`$ (83)
Coeficientul de reflexie cerut este prin urmare
$`R={\displaystyle \frac{|\psi _{refl}|^2}{|\psi _{inc}|^2}}|_{x=0}={\displaystyle \frac{|B|^2}{|A|^2}}={\displaystyle \frac{m^2V_0^2}{m^2V_0^2+\mathrm{}^4k^2}}.`$ (84)
Dacฤ potenลฃialul este extrem de puternic ($`V_0\mathrm{}`$) se vede cฤ $`R1`$, adicฤ unda este reflectatฤ รฎn totalitate.
Coeficientul de transmisie, pe de altฤ parte, este
$`T={\displaystyle \frac{|\psi _{trans}|^2}{|\psi _{inc}|^2}}|_{x=0}={\displaystyle \frac{|C|^2}{|A|^2}}={\displaystyle \frac{\mathrm{}^4k^2}{m^2V_0^2+\mathrm{}^4k^2}}.`$ (85)
Dacฤ potenลฃialul este foarte puternic ($`V_0\mathrm{}`$) atunci $`T0`$, adicฤ , unda transmisฤ cade rapid de cealaltฤ parte a potenลฃialului.
Evident, $`R+T=1`$ cum era de aลteptat.
### Problema 2.2: Particulฤ รฎntr-o groapฤ de potenลฃial finitฤ 1D
Sฤ se rezolve ecuaลฃia Schrรถdinger unidimensionalฤ pentru o groapฤ de potenลฃial finitฤ descrisฤ prin condiลฃiile
$$V(x)=\{\begin{array}{cc}V_0\hfill & \text{dacฤ }|x|a\hfill \\ 0\hfill & \text{dacฤ }|x|>a\text{ .}\hfill \end{array}$$
Sฤ se considere numai stฤrile legate ($`E<0`$).
Soluลฃie.
a) Funcลฃia de undฤ pentru $`|x|<a`$ ลi $`|x|>a`$.
Ecuaลฃia Schrรถdinger corespunzฤtoare este
$$\frac{\mathrm{}^2}{2m}\psi ^{^{\prime \prime }}(x)+V(x)\psi (x)=E\psi (x).$$
(86)
Definim
$$q^2=\frac{2mE}{\mathrm{}^2},k^2=\frac{2m(E+V_0)}{\mathrm{}^2}$$
(87)
ลi obลฃinem:
$`1)\mathrm{pentru}\mathrm{x}<\mathrm{a}:\psi ^{^{\prime \prime }}_1(x)q^2\psi _1`$ $`=`$ $`0,\psi _1=A_1e^{qx}+B_1e^{qx};`$
$`2)\mathrm{pentru}\mathrm{a}\mathrm{x}\mathrm{a}:\psi ^{^{\prime \prime }}_2(x)+k^2\psi _2`$ $`=`$ $`0,\psi _2=A_2\mathrm{cos}(kx)+B_2\mathrm{sin}(kx);`$
$`3)\mathrm{pentru}\mathrm{x}>\mathrm{a}:\psi ^{^{\prime \prime }}_3(x)q^2\psi _3`$ $`=`$ $`0,\psi _3=B_3e^{qx}+B_3e^{qx}.`$
b) Formularea condiลฃiilor de frontierฤ .
Normalizarea stฤrilor legate cere ca soluลฃia sฤ fie zero la infinit. Aceasta รฎnseamnฤ cฤ $`B_1=A_3=0`$. รn plus, $`\psi (x)`$ trebuie sฤ fie continuu diferenลฃiabilฤ . Toate soluลฃiile particulare sunt fixate รฎn aลa fel รฎncรฎt $`\psi `$ precum ลi prima sa derivatฤ $`\psi ^{}`$ sunt continue รฎn acea valoare a lui x corespunzรฎnd frontierei รฎntre zona interioarฤ ลi cea exterioarฤ . A doua derivatฤ $`\psi ^{\prime \prime }`$ conลฃine saltul (discontinuitatea) impus de cฤtre potenลฃialul particular de tip โcutieโ al acestei ecuaลฃii Schrรถdinger. Toate acestea ne conduc la
$`\psi _1(a)`$ $`=`$ $`\psi _2(a),\psi _2(a)=\psi _3(a),`$
$`\psi _1^{}(a)`$ $`=`$ $`\psi _2^{}(a),\psi _2^{}(a)=\psi _3^{}(a).`$ (88)
c) Ecuaลฃiile de valori proprii.
Din (88) obลฃinem patru ecuaลฃii lineare ลi omogene pentru coeficienลฃii $`A_1`$, $`A_2`$, $`B_2`$ ลi $`B_3`$:
$`A_1e^{qa}`$ $`=`$ $`A_2\mathrm{cos}(ka)B_2\mathrm{sin}(ka),`$
$`qA_1e^{qa}`$ $`=`$ $`A_2k\mathrm{sin}(ka)+B_2k\mathrm{cos}(ka),`$
$`B_3e^{qa}`$ $`=`$ $`A_2\mathrm{cos}(ka)+B_2\mathrm{sin}(ka),`$
$`qB_3e^{qa}`$ $`=`$ $`A_2k\mathrm{sin}(ka)+B_2k\mathrm{cos}(ka).`$ (89)
Adunรฎnd ลi scฤzรฎnd obลฃinem un sistem de ecuaลฃii mai uลor de rezolvat:
$`(A_1+B_3)e^{qa}`$ $`=`$ $`2A_2\mathrm{cos}(ka)`$
$`q(A_1+B_3)e^{qa}`$ $`=`$ $`2A_2k\mathrm{sin}(ka)`$
$`(A_1B_3)e^{qa}`$ $`=`$ $`2B_2\mathrm{sin}(ka)`$
$`q(A_1B_3)e^{qa}`$ $`=`$ $`2B_2k\mathrm{cos}(ka).`$ (90)
Dacฤ se presupune cฤ $`A_1+B_30`$ ลi $`A_20`$, primele douฤ ecuaลฃii dau
$$q=k\mathrm{tan}(ka).$$
(91)
care pus รฎn ultimele douฤ dฤ
$$A_1=B_3;B_2=0.$$
(92)
De aici, ca rezultat, obลฃinem o soluลฃie simetricฤ , $`\psi (x)=\psi (x)`$, sau de paritate pozitivฤ .
Un calcul practic identic ne duce pentru $`A_1B_30`$ ลi pentru $`B_20`$ la
$$q=k\mathrm{cot}(ka)yA_1=B_3;A_2=0.$$
(93)
Funcลฃia de undฤ astfel obลฃinutฤ este antisimetricฤ , corespunzรฎnd unei paritฤลฃi negative.
d) Soluลฃie calitativฤ a problemei de valori proprii.
Ecuaลฃia care leagฤ $`q`$ ลi $`k`$, pe care am obลฃinut-o deja, dฤ condiลฃii pentru autovalorile de energie. Folosind forma scurtฤ
$$\xi =ka,\eta =qa,$$
(94)
obลฃinem din definiลฃia (87)
$$\xi ^2+\eta ^2=\frac{2mV_0a^2}{\mathrm{}^2}=r^2.$$
(95)
Pe de altฤ parte, folosind (91) ลi (93) obลฃinem ecuaลฃiile
$`\eta =\xi \mathrm{tan}(\xi ),\eta =\xi \mathrm{cot}(\xi ).`$
Astfel autovalorile de energie cฤutate pot fi obลฃinute construind intersecลฃia acestor douฤ curbe cu cercul definit de (95), รฎn planul $`\xi `$-$`\eta `$ (vezi figura 2.6).
Cel puลฃin o soluลฃie existฤ pentru valori arbitrare ale parametrului $`V_0`$, รฎn cazul paritฤลฃii pozitive, pentru cฤ funcลฃia tangentฤ intersecteazฤ originea. Pentru paritatea negativฤ , raza cercului trebuie sฤ fie mai mare decรฎt o anumitฤ valoare minimฤ astfel cฤ cele douฤ curbe sฤ se poatฤ intersecta. Potenลฃialul trebuie sฤ aibฤ o anumitฤ adรฎncime relaลฃionatฤ cu o scalฤ spaลฃialฤ datฤ $`a`$ ลi o scalฤ de masฤ datฤ $`m`$, pentru a permite o soluลฃie de paritate negativฤ . Numฤrul de nivele de energie creลte cu $`V_0`$, $`a`$ ลi masa $`m`$. Pentru cazul รฎn care $`mVa^2\mathrm{}`$, intersecลฃiile se aflฤ din
$`\mathrm{tan}(ka)`$ $`=`$ $`\mathrm{}ka={\displaystyle \frac{2n1}{2}}\pi ,`$
$`\mathrm{cot}(ka)`$ $`=`$ $`\mathrm{}ka=n\pi ,`$ (96)
unde $`n=1,2,3,\mathrm{}`$
sau combinรฎnd:
$$k(2a)=n\pi .$$
(97)
Pentru spectrul de energie acest lucru รฎnseamnฤ cฤ
$$E_n=\frac{\mathrm{}^2}{2m}(\frac{n\pi }{2a})^2V_0.$$
(98)
Lฤrgind groapa de potenลฃial ลi/sau masa particulei $`m`$, diferenลฃa รฎntre douฤ autovalori de energie vecine va descreลte. Nivelul cel mai de jos ($`n=1`$) nu este localizat รฎn $`V_0`$, ci un pic mai sus. Aceastฤ diferenลฃฤ se numeลte energia de punct zero.
e) Formele funcลฃiilor de undฤ se aratฤ รฎn figura 2.7 pentru soluลฃiile discutate .
### Problema 2.3: Particulฤ รฎn groapฤ rectangularฤ 1D infinitฤ
Sฤ se rezolve ecuaลฃia Schrรถdinger unidimensionalฤ pentru o particulฤ care se aflฤ intr-o groapฤ de potenลฃial infinitฤ descrisฤ de:
$$V(x)=\{\begin{array}{cc}0\hfill & \text{pentru }x^{}<x<x^{}+2a\hfill \\ \mathrm{}\hfill & \text{pentru }x^{}x\mathrm{o}xx^{}+2a\text{.}\hfill \end{array}$$
Soluลฃia รฎn formฤ generalฤ este
$$\psi (x)=A\mathrm{sin}(kx)+B\mathrm{cos}(kx)$$
(99)
unde
$$k=\sqrt{\frac{2mE}{\mathrm{}^2}}.$$
(100)
Cum $`\psi `$ trebuie sฤ รฎndeplineascฤ $`\psi (x^{})=\psi (x^{}+2a)=0`$, se obลฃine:
$`A\mathrm{sin}(kx^{})+B\mathrm{cos}(kx^{})=0`$ (101)
$`A\mathrm{sin}[k(x^{}+2a)]+B\mathrm{cos}[k(x^{}+2a)]=0.`$ (102)
Multiplicรฎnd (101) cu $`\mathrm{sin}[k(x^{}+2a)]`$ ลi (102) cu $`\mathrm{sin}(kx^{})`$ ลi รฎn continuare scฤzรฎnd ultimul rezultat din primul obลฃinem:
$$B[\mathrm{cos}(kx^{})\mathrm{sin}[k(x^{}+2a)]\mathrm{cos}[k(x^{}+2a)]\mathrm{sin}(kx^{})]=0.$$
(103)
sau prin intermediul unei identitฤลฃi trigonometrice:
$$B\mathrm{sin}(2ak)=0$$
(104)
Multiplicรฎnd (101) cu $`\mathrm{cos}[k(x^{}+2a)]`$ ลi scฤzรฎnd (102) multiplicatฤ cu $`\mathrm{cos}(kx^{})`$, se obลฃine:
$$A[\mathrm{sin}(kx^{})\mathrm{cos}[k(x^{}+2a)]\mathrm{sin}[k(x^{}+2a)]\mathrm{cos}(kx^{})]=0,$$
(105)
sau folosind aceeaลi identitate trigonometricฤ :
$$A\mathrm{sin}[k(2ak)]=A\mathrm{sin}[k(2ak)]=0.$$
(106)
Cum nu se ลฃine cont de soluลฃia trivialฤ $`\psi =0`$, atunci pe baza ecuaลฃiilor (104) ลi (106) trebuie ca $`\mathrm{sin}(2ak)=0`$, care are loc numai dacฤ $`2ak=n\pi `$, cu $`n`$ un รฎntreg. Conform celor anterioare $`k=n\pi /2a`$ ลi cum $`k^2=2mE/\mathrm{}^2`$ atunci rezultฤ cฤ autovalorile sunt date de cฤtre expresia:
$$E=\frac{\mathrm{}^2\pi ^2n^2}{8a^2m}.$$
(107)
Energia este cuantizatฤ pentru cฤ numai pentru fiecare $`k_n=n\pi /2a`$ le corespunde o energie $`E_n=[n^2/2m][\pi \mathrm{}/2a]^2`$.
Soluลฃia รฎn forma generalฤ :
$$\psi _n=A\mathrm{sin}(\frac{n\pi x}{2a})+B\mathrm{cos}(\frac{n\pi x}{2a}).$$
(108)
se poate normaliza:
$$1=_x^{}^{x^{}+2a}\psi \psi ^{}๐x=a(A^2+B^2)$$
(109)
ceea ce ne conduce la:
$$A=\pm \sqrt{1/aB^2}.$$
(110)
Substituind aceastฤ valoare a lui $`A`$ รฎn (101) se obลฃine:
$$B=\frac{1}{\sqrt{a}}\mathrm{sin}(\frac{n\pi x^{}}{2a}).$$
(111)
Substituind acum aceasta valoare a lui $`B`$ รฎn (110) se obลฃine:
$$A=\pm \frac{1}{\sqrt{a}}\mathrm{cos}(\frac{n\pi x^{}}{2a})$$
(112)
Folosind semnele superioare pentru $`A`$ ลi $`B`$, prin substituirea valorilor acestora รฎn (108) se obลฃine:
$$\psi _n=\frac{1}{\sqrt{a}}\mathrm{sin}(\frac{n\pi }{2a})(xx^{}).$$
(113)
Utilizรฎnd semnele inferioare pentru A ลi B se obลฃine:
$$\psi _n=\frac{1}{\sqrt{a}}\mathrm{sin}(\frac{n\pi }{2a})(xx^{}).$$
(114)
3. MOMENT CINETIC ลI SPIN
## Introducere
Din Mecanica Clasicฤ se ลtie cฤ , momentul cinetic $`๐ฅ`$ pentru particulele macroscopice este dat de
$$๐ฅ=๐ซ\times ๐ฉ,$$
(1)
unde $`๐ซ`$ ลi $`๐ฉ`$ sunt respectiv vectorul de poziลฃie ลi impulsul lineal.
Totuลi, รฎn Mecanica Cuanticฤ , existฤ operatori de tip moment cinetic (OTMC) care nu obligatoriu se exprimฤ (numai) prin operatorii de coordonatฤ $`\widehat{x}_j`$ ลi impuls $`\widehat{p}_k`$ acลฃionรฎnd (numai) asupra funcลฃiilor proprii de coordonate. Prin urmare, este foarte important sฤ se stabileascฤ , mai รฎntรฎi de toate, relaลฃii de comutare generale pentru componentele OTMC.
รn Mecanica Cuanticฤ $`๐ฅ`$ se reprezintฤ prin operatorul
$$๐ฅ=i\mathrm{}๐ซ\times ,$$
(2)
ale cฤrui componente sunt operatori care satisfac urmฤtoarele reguli de comutare
$$[l_x,l_y]=il_z,[l_y,l_z]=il_x,[l_z,l_x]=il_y,$$
(3)
ลi รฎn plus, fiecare dintre ele comutฤ cu pฤtratul momentului cinetic, adicฤ
$$l^2=l_x^2+l_y^2+l_z^2,[l_i,l^2]=0,i=1,2,3.$$
(4)
Aceste relaลฃii, รฎn afarฤ de a fi corecte pentru momentul cinetic, se satisfac pentru importanta clasฤ OTMC a operatorilor de spin, care sunt lipsiลฃi de analogi รฎn mecanica clasicฤ .
Aceste relaลฃii de comutare sunt bazice pentru a obลฃine spectrul operatorilor menลฃionaลฃi, precum ลi reprezentฤrile lor diferenลฃiale.
## Momentul cinetic orbital
Pentru un punct oarecare al unui spaลฃiu fix (SF), se poate introduce o funcลฃie $`\psi (x,y,z)`$, pentru care, sฤ considerฤm douฤ sisteme carteziene $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$, unde $`\mathrm{\Sigma }^{}`$ se obลฃine prin rotaลฃia axei $`z`$ a lui $`\mathrm{\Sigma }`$.
รn cazul general un SF se referฤ la un sistem de coordonate diferite de $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$.
Acum, sฤ comparฤm valorile lui $`\psi `$ รฎn douฤ puncte ale SF cu aceleaลi coordonate (x,y,z) รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$, ceea ce este echivalent cu rotaลฃia vectorialฤ
$$\psi (x^{},y^{},z^{})=R\psi (x,y,z)$$
(5)
unde $`R`$ este matricea de rotaลฃie รฎn R<sup>3</sup>
$$\left(\begin{array}{c}x^{}\\ y^{}\\ z^{}\end{array}\right)=\left(\begin{array}{ccc}\mathrm{cos}\varphi & \mathrm{sin}\varphi & 0\\ \mathrm{sin}\varphi & \mathrm{cos}\varphi & 0\\ 0& 0& z\end{array}\right)\left(\begin{array}{c}x\\ y\\ z\end{array}\right),$$
(6)
atunci
$$R\psi (x,y,z)=\psi (x\mathrm{cos}\varphi y\mathrm{sin}\varphi ,x\mathrm{sin}\varphi +y\mathrm{cos}\varphi ,z).$$
(7)
Pe de altฤ parte este important de amintit cฤ funcลฃiile de undฤ nu depind de sistemul de coordonate ลi cฤ trasformarea la rotaลฃii รฎn cadrul SF se face cu ajutorul operatorilor unitari ลi deci pentru a stabili forma operatorului unitar $`U^{}(\varphi )`$ care trece $`\psi `$ รฎn $`\psi ^{}`$, se considerฤ o rotaลฃie infinitezimalฤ $`d\varphi `$, menลฃinรฎnd numai termenii lineari รฎn $`d\varphi `$ atunci cรฎnd se face dezvoltarea รฎn serie Taylor a lui $`\psi ^{}`$ รฎn jurul punctului $`x`$
$`\psi (x^{},y^{},z^{})`$ $``$ $`\psi (x+yd\varphi ,xd\varphi +y,z),`$ (8)
$``$ $`\psi (x,y,z)+d\varphi \left(y{\displaystyle \frac{\psi }{x}}x{\displaystyle \frac{\psi }{y}}\right),`$
$``$ $`(1id\varphi l_z)\psi (x,y,z),`$
unde am folosit notaลฃia<sup>1</sup><sup>1</sup>1Demostraลฃia lui (8) se prezintฤ ca problema 3.1
$$l_z=\mathrm{}^1(\widehat{x}\widehat{p}_y\widehat{y}\widehat{p}_x),$$
(9)
care, dupฤ cum se va vedea mai tรฎrziu, corespunde operatorului de proiecลฃie รฎn $`z`$ al momentului cinetic de acord cu definiลฃia din (2) ลi adฤugรฎnd factorul $`\mathrm{}^1`$, astfel cฤ rotaลฃiile de unghi $`\varphi `$ finit se pot reprezenta ca exponenลฃiale de tipul:
$$\psi (x^{},y^{},z)=e^{il_z\varphi }\psi (x,y,z),$$
(10)
unde
$$\widehat{U}^{}(\varphi )=e^{il_z\varphi }.$$
(11)
Pentru a reafirma conceptul de rotaลฃie, sฤ รฎl considerฤm รฎntr-un tratament mai general cu ajutorul operatorului vectorial $`\widehat{\stackrel{}{A}}`$ care acลฃioneazฤ asupra lui $`\psi `$, presupunรฎnd cฤ $`\widehat{A}_x`$, $`\widehat{A}_y`$ $`\widehat{A}_z`$ au aceeaลi formฤ รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$, adicฤ , valorile medii ale lui $`\widehat{\stackrel{}{A}}`$ calculate รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$ trebuie sฤ fie egale cรฎnd se vฤd din SF:
$`{\displaystyle \psi ^{}(\stackrel{}{r}^{})(\widehat{A}_x\widehat{ฤฑ}^{}+\widehat{A}_y\widehat{ศท}^{}+\widehat{A}_z\widehat{k}^{})\psi ^{}(\stackrel{}{r}^{})๐\stackrel{}{r}}`$
$`={\displaystyle \psi ^{}(\stackrel{}{r})(\widehat{A}_x\widehat{ฤฑ}+\widehat{A}_y\widehat{ศท}+\widehat{A}_z\widehat{k})\psi ^{}(\stackrel{}{r})๐\stackrel{}{r}},`$ (12)
unde
$$\widehat{ฤฑ}^{}=\widehat{ฤฑ}\mathrm{cos}\varphi +\widehat{ศท}\mathrm{sin}\varphi ,\widehat{ศท}^{}=\widehat{ฤฑ}\mathrm{sin}\varphi +\widehat{ศท}\mathrm{cos}\varphi ,\widehat{k}^{}=\widehat{k}.$$
(13)
Prin urmare, dacฤ vom combina (10), (12) ลi (13) obลฃinem
$`e^{il_z\varphi }\widehat{A}_xe^{il_z\varphi }`$ $`=`$ $`\widehat{A}_x\mathrm{cos}\varphi \widehat{A}_y\mathrm{sin}\varphi ,`$
$`e^{il_z\varphi }\widehat{A}_ye^{il_z\varphi }`$ $`=`$ $`\widehat{A}_x\mathrm{sin}\varphi \widehat{A}_y\mathrm{cos}\varphi ,`$
$`e^{il_z\varphi }\widehat{A}_ze^{il_z\varphi }`$ $`=`$ $`\widehat{A}_z.`$ (14)
Din nou considerรฎnd rotaลฃii infinitezimale ลi dezvoltรฎnd pฤrลฃile din stรฎnga รฎn (14) se pot determina relaลฃiile de comutare ale lui $`\widehat{A}_x`$, $`\widehat{A}_y`$ ลi $`\widehat{A}_z`$ cu $`\widehat{l}_z`$
$$[l_z,A_x]=iA_y,[l_z,A_y]=iA_x,[l_z,A_z]=0,$$
(15)
ลi รฎn acelaลi mod pentru $`l_x`$ ลi $`l_y`$.
Condiลฃiile bazice pentru a obลฃine aceste relaลฃii de comutare sunt
* FP se transformฤ ca รฎn (7) cรฎnd $`\mathrm{\Sigma }\mathrm{\Sigma }^{}`$.
* Componentele $`\widehat{A}_x`$, $`\widehat{A}_y`$, $`\widehat{A}_z`$ au aceeaลi formฤ รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$.
* Vectorii valorilor medii ale lui $`\widehat{A}`$ รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$ coincid (sunt egale) pentru un observator din SF.
Deasemenea se poate folosi altฤ reprezentare รฎn care $`\psi (x,y,z)`$ nu se schimbฤ cรฎnd $`\mathrm{\Sigma }\mathrm{\Sigma }^{}`$ ลi operatorii vectoriali se transformฤ ca orice vectori. Pentru a trece la o astfel de representare cรฎnd $`\varphi `$ se roteลte รฎn jurul lui $`z`$ se foloseลte operatorul $`\widehat{U}(\varphi )`$, adicฤ
$$e^{il_z\varphi }\psi ^{}(x,y,z)=\psi (x,y,z),$$
(16)
ลi deci
$$e^{il_z\varphi }\widehat{\stackrel{}{A}}e^{il_z\varphi }=\widehat{\stackrel{}{A}}.$$
(17)
Utilizรฎnd relaลฃiile date รฎn (14) obลฃinem
$`\widehat{A}_x^{}`$ $`=`$ $`\widehat{A}_x\mathrm{cos}\varphi +\widehat{A}_y\mathrm{sin}\varphi =e^{il_z\varphi }\widehat{A}_xe^{il_z\varphi },`$
$`\widehat{A}_y^{}`$ $`=`$ $`\widehat{A}_x\mathrm{sin}\varphi +\widehat{A}_y\mathrm{cos}\varphi =e^{il_z\varphi }\widehat{A}_ye^{il_z\varphi },`$
$`\widehat{A}_z^{}`$ $`=`$ $`e^{il_z\varphi }\widehat{A}_ze^{il_z\varphi }.`$ (18)
Pentru cฤ transformฤrile noii reprezentฤri se fac cu ajutorul operatorilor unitari, relaลฃiile de comutare nu se schimbฤ .
### Observaลฃii
* Operatorul $`\widehat{A}^2`$ este invariant la rotaลฃii, adicฤ
$$e^{il_z\varphi }\widehat{A}^2e^{il_z\varphi }=\widehat{A}^2=\widehat{A}^2.$$
(19)
* Rezultฤ cฤ
$$[\widehat{l}_i,\widehat{A}^2]=0.$$
(20)
* Dacฤ operatorul Hamiltonian este de forma
$$\widehat{H}=\frac{1}{2m}\widehat{p}^2+U(|\stackrel{}{r}|),$$
(21)
atunci se menลฃine invariant la rotaลฃii produse รฎn oricare axฤ care trece prin originea de coordonate
$$[\widehat{l}_i,\widehat{H}]=0,$$
(22)
unde $`\widehat{l}_i`$ sunt integrale de miลcare.
### Definiลฃie
Dacฤ $`\widehat{A}_i`$ sunt componentele unui operator vectorial care acลฃioneazฤ asupra unei funcลฃii de undฤ dependentฤ numai de coordonate ลi dacฤ existฤ operatori $`\widehat{l}_i`$ care satisfac urmฤtoarele relaลฃii de comutare:
$$[\widehat{l}_i,\widehat{A}_j]=i\epsilon _{ijk}\widehat{A}_k,[\widehat{l}_i,\widehat{l}_j]=i\epsilon _{ijk}\widehat{l}_k,$$
(23)
atunci, $`\widehat{l}_i`$ se numesc componentele operatorului moment cinetic (orbital) ลi putem sฤ deducem pe baza lui (20) ลi (23) cฤ
$$[\widehat{l}_i,\widehat{l}^2]=0.$$
(24)
Prin urmare, cele trei componente operatoriale asociate cu componentele unui moment cinetic clasic arbitrar satisfac relaลฃii de comutare de tipul (23), (24). Mai mult, se poate arฤta cฤ aceste relaลฃii conduc la proprietฤลฃi geometrice specifice ale rotaลฃiilor รฎntr-un spaลฃiu euclidean tridimensional. Aceasta are loc dacฤ adoptฤm un punct de vedere mai general ลi definim un operator moment unghiular $`๐`$ (nu mai punem simbolul de operator) ca orice set de trei observabile $`J_x`$, $`J_y`$ ลi $`J_z`$ care รฎndeplinesc relaลฃiile de comutare:
$$[J_i,J_j]=i\epsilon _{ijk}J_k.$$
(25)
รn plus, sฤ introducem operatorul
$$๐^2=J_x^2+J_y^2+J_z^2,$$
(26)
pฤtratul scalar al momentului unghiular $`๐`$. Acest operator este hermitic pentru cฤ $`J_x`$, $`J_y`$ ลi $`J_z`$ sunt hermitici ลi se aratฤ uลor cฤ $`๐^\mathrm{๐}`$ comutฤ cu cele trei componente ale lui $`๐`$
$$[๐^2,J_i]=0.$$
(27)
Deoarece $`๐^\mathrm{๐}`$ comutฤ cu fiecare dintre componente rezultฤ cฤ existฤ un sistem complet de FP, respectiv
$$๐^\mathrm{๐}\psi _{\gamma \mu }=f(\gamma ^2)\psi _{\gamma \mu },J_i\psi _{\gamma \mu }=g(\mu )\psi _{\gamma \mu },$$
(28)
unde, aลa cum se va arฤta รฎn continuare, FP-urile depind de doi subindici, care se vor determina odatฤ cu forma funcลฃiilor $`f(\gamma )`$ ลi $`g(\mu )`$. Operatorii $`J_i`$ ลi $`J_k`$ $`(ik)`$ nu comutฤ (รฎntre ei), adicฤ , nu au FP รฎn comun. Din motive atรฎt fizice cรฎt ลi matematice, suntem interesaลฃi sฤ determinฤm funcลฃiile proprii comune ale lui $`๐^\mathrm{๐}`$ ลi $`J_z`$, adicฤ vom lua $`i=z`$ รฎn (28).
รn loc de a folosi componentele $`J_x`$ ลi $`J_y`$ ale momentului unghiular $`๐`$, este mai convenabil sฤ se introducฤ urmฤtoarele combinaลฃii lineare
$$J_+=J_x+iJ_y,J_{}=J_xiJ_y.$$
(29)
Aceลti operatori nu sunt hermitici, aลa cum sunt operatorii $`a`$ ลi $`a^{}`$ ai oscilatorului armonic (vezi capitolul 5), sunt numai adjuncลฃi. Se pot deduce uลor urmฤtoarele proprietฤลฃi
$$[J_z,J_\pm ]=\pm J_\pm ,[J_+,J_{}]=2J_z,$$
(30)
$$[J^2,J_+]=[J^2,J_{}]=[J^2,J_z]=0.$$
(31)
$$J_z(J_\pm \psi _{\gamma \mu })=\{J_\pm J_z+[J_z,J_\pm ]\}\psi _{\gamma \mu }=(\mu \pm 1)(J_\pm \psi _{\gamma \mu }).$$
(32)
Prin urmare $`J_\pm \psi _{\gamma \mu }`$ sunt FP ale lui $`J_z`$ corespunzฤtoare autovalorilor $`\mu \pm 1`$, adicฤ funcลฃiile respective sunt identice pรฎnฤ la factorii constaลฃi (de determinat) $`\alpha _\mu `$ ลi $`\beta _\mu `$:
$`J_+\psi _{\gamma \mu 1}`$ $`=`$ $`\alpha _\mu \psi _{\gamma \mu },`$
$`J_{}\psi _{\gamma \mu }`$ $`=`$ $`\beta _\mu \psi _{\gamma \mu 1}.`$ (33)
Pe de altฤ parte
$$\alpha _\mu ^{}=(J_+\psi _{\gamma \mu 1},\psi _{\gamma \mu })=(\psi _{\gamma \mu 1}J_{}\psi _{\gamma \mu })=\beta _\mu ,$$
(34)
astfel cฤ , luรฎnd o fazฤ de tipul $`e^{ia}`$ (cu $`a`$ real) pentru funcลฃia $`\psi _{\gamma \mu }`$ se poate face $`\alpha _\mu `$ real ลi egal cu $`\beta _\mu `$, ceea ce รฎnseamnฤ
$$J_+\psi _{\gamma ,\mu 1}=\alpha \mu \psi _{\gamma \mu },J_{}\psi _{\gamma \mu }=\alpha \mu \psi _{\gamma ,\mu 1},$$
(35)
ลi deci
$`\gamma `$ $`=`$ $`(\psi _{\gamma \mu },[J_x^2+J_y^2+J_z^2]\psi _{\gamma \mu })=\mu ^2+a+b,`$
$`a`$ $`=`$ $`(\psi _{\gamma \mu },J_x^2\psi _{\gamma \mu })=(J_x\psi _{\gamma \mu },J_x\psi _{\gamma \mu })0,`$
$`b`$ $`=`$ $`(\psi _{\gamma \mu },J_y^2\psi _{\gamma \mu })=(J_y\psi _{\gamma \mu },J_y\psi _{\gamma \mu })0.`$ (36)
Constanta de normalizare a oricฤrei funcลฃii de undฤ nu este negativฤ , ceea ce implicฤ
$$\gamma \mu ^2,$$
(37)
pentru $`\gamma `$ fix, valoarea lui $`\mu `$ are limite atรฎt superioarฤ cรฎt ลi inferioarฤ (adicฤ ia valori รฎntr-un interval finit).
Fie $`\mathrm{\Lambda }`$ ลi $`\lambda `$ aceste limite (superioarฤลi inferiarฤ de $`\mu `$) pentru un $`\gamma `$ dat
$$J_+\psi _{\gamma \mathrm{\Lambda }}=0,J_{}\psi _{\gamma \lambda }=0.$$
(38)
Utilizรฎnd urmฤtoarele identitฤลฃi operatoriale
$`J_{}J_+`$ $`=`$ $`๐^\mathrm{๐}J_z^2+J_z=๐^\mathrm{๐}J_z(J_z1),`$
$`J_+J_{}`$ $`=`$ $`๐^\mathrm{๐}J_z^2+J_z=๐^\mathrm{๐}J_z(J_z+1),`$ (39)
acลฃionรฎnd asupra lui $`\psi _{\gamma \mathrm{\Lambda }}`$ ลi $`\psi _{\gamma \lambda }`$ se obลฃine
$`\gamma \mathrm{\Lambda }^2\mathrm{\Lambda }`$ $`=`$ $`0,`$
$`\gamma \lambda ^2+\lambda `$ $`=`$ $`0,`$
$`(\lambda \lambda +1)(\lambda +\lambda )`$ $`=`$ $`0.`$ (40)
รn plus
$$\mathrm{\Lambda }\lambda \mathrm{\Lambda }=\lambda =J\gamma =J(J+1).$$
(41)
Pentru un $`\gamma `$ dat (fix) proiecลฃia momentului $`\mu `$ ia $`2J+1`$ valori care diferฤ printr-o unitate รฎntre $`J`$ ลi $`J`$. De aceea, diferenลฃa $`\mathrm{\Lambda }\lambda =2J`$ trebuie sฤ fie un numฤr รฎntreg ลi deci autovalorile lui $`J_z`$ numerotate prin $`m`$ sunt รฎntregi
$$m=k,k=0,\pm 1,\pm 2,\mathrm{},$$
(42)
sau semiรฎntregi
$$m=k+\frac{1}{2},k=0,\pm 1,\pm 2,\mathrm{}.$$
(43)
O stare de $`\gamma =J(J+1)`$ datฤ indicฤ o degenerare de grad $`g=2J+1`$ faลฃฤ de autovalorile $`m`$ (aceasta pentru cฤ $`J_i,J_k`$ comutฤ cu $`J^2`$ dar nu comutฤ รฎntre ele).
Prin โstare de moment unghiular $`J`$โ se รฎnลฃelege รฎn majoritatea cazurilor, o stare cu $`\gamma =J(J+1)`$ care are valoarea maximฤ a momentului proiectat, respectiv $`J`$. Este comun sฤ se noteze aceste stฤri cu $`\psi _{jm}`$ sau de ket Dirac $`|jm`$.
Sฤ obลฃinem acum elementele de matrice ale lui $`J_x,J_y`$ รฎn reprezentarea รฎn care $`J^2`$ ลi $`J_z`$ sunt diagonale. รn acest caz, din (35) ลi (39) se obลฃine cฤ
$`J_{}J_+\psi _{jm1}=\alpha _mJ_{}\psi _{jm}=\alpha _m\psi _{jm1},`$
$`J(J+1)(m1)^2(m1)=\alpha _m^2,`$
$`\alpha _m=\sqrt{(J+m)(Jm+1)}.`$ (44)
Combinรฎnd (44) cu (35) se obลฃine
$$J_+\psi _{jm1}=\sqrt{(J+m)(Jm+1)}\psi _{jm},$$
(45)
rezultฤ cฤ elementul de matrice al lui $`J_+`$ este
$$jm|J_+|jm1=\sqrt{(J+m)(Jm+1)}\delta _{nm},$$
(46)
ลi analog
$$jn|J_{}|jm=\sqrt{(J+m)(Jm+1)}\delta _{nm1}.$$
(47)
รn sfรฎrลit, din definiลฃiile (29) pentru $`J_+,J_{}`$ se obลฃine uลor
$`jm|J_x|jm1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{(J+m)(Jm+1)},`$
$`jm|J_y|jm1`$ $`=`$ $`{\displaystyle \frac{i}{2}}\sqrt{(J+m)(Jm+1)}.`$ (48)
### Concluzii parลฃiale
* Proprietฤลฃi ale autovalorilor lui $`๐`$ ลi $`J_z`$
Dacฤ $`j(j+1)\mathrm{}^2`$ ลi $`m\mathrm{}`$ sunt autovalori ale lui $`๐`$ ลi $`J_z`$ asociate cu autovectorii $`|kjm`$, atunci $`j`$ ลi $`m`$ satisfac inegalitatea
$$jmj.$$
* Proprietฤลฃi ale vectorului $`J_{}|kjm`$
Fie $`|kjm`$ un eigenvector al lui $`๐^\mathrm{๐}`$ ลi $`J_z`$ cu autovalorile $`j(j+1)\mathrm{}^2`$ ลi $`m\mathrm{}`$
+ (i) Dacฤ $`m=j`$ atunci $`J_{}|kjj=0`$.
+ (ii) Dacฤ $`m>j`$ atunci $`J_{}|kjm`$ este un vector propriu nenul al lui $`J^2`$ ลi $`J_z`$ cu autovalorile $`j(j+1)\mathrm{}^2`$ ลi $`(m1)\mathrm{}`$.
* Proprietฤลฃi ale vectorului $`J_+|kjm`$
Fie $`|kjm`$ un vector (ket) propriu al lui $`๐^\mathrm{๐}`$ ลi $`J_z`$ cu autovalorile $`j(j+1)\mathrm{}`$ ลi $`m\mathrm{}`$
+ Dacฤ $`m=j`$ atunci $`J_+|kjm=0.`$
+ Dacฤ $`m<j`$ atunci $`J_+|kjm`$ este un vector nenul al lui $`๐^\mathrm{๐}`$ ลi $`J_z`$ cu autovalorile $`j(j+1)\mathrm{}^2`$ ลi $`(m+1)\mathrm{}`$
* Consecinลฃe ale proprietฤลฃilor anterioare
$`J_z|kjm`$ $`=`$ $`m\mathrm{}|kjm,`$
$`J_+|kjm`$ $`=`$ $`m\mathrm{}\sqrt{j(j+1)m(m+1)}|kjm+1,`$
$`J_{}|kjm`$ $`=`$ $`m\mathrm{}\sqrt{j(j+1)m(m1)}|kjm+1.`$
## Aplicaลฃii ale momentului cinetic orbital
Am considerat pรฎnฤ acum proprietฤลฃile momentului cinetic derivate numai pe baza relaลฃiilor de comutare. Sฤ ne รฎntoarcem la momentul cinetic $`๐ฅ`$ al unei particule fฤrฤ rotaลฃie intrinsecฤ ลi sฤ vedem cum se aplicฤ teoria din secลฃiunea anterioarฤ la cazul particular important
$$[\widehat{l}_i,\widehat{p}_j]=i\epsilon _{ijk}\widehat{p}_k.$$
(49)
Mai รฎntรฎi, $`\widehat{l}_z`$ ลi $`\widehat{p}_j`$ au un sistem comun de funcลฃii proprii. Pe de altฤ parte, Hamiltonianul unei particule libere
$$\widehat{H}=\left(\frac{\widehat{\stackrel{}{p}}}{\sqrt{2m}}\right)^2,$$
fiind pฤtratul unui operator vectorial are acelaลi sistem complet de FP cu $`\widehat{l^2}`$ ลi $`\widehat{l}_z`$. รn plus, aceasta implicฤ cฤ particula liberฤ se poate gฤsi รฎntr-o stare cu $`E`$, $`l`$, $`m`$ bine determinate.
### Valori proprii ลi funcลฃii proprii ale lui $`๐ฅ^\mathrm{๐}`$ ลi $`๐_๐ณ`$
Este mai convenabil sฤ se lucreze cu coordonate sferice (sau polare), pentru cฤ , aลa cum vom vedea, diferiลฃi operatori de moment cinetic acลฃioneazฤ numai asupra variabilelor unghiulare $`\theta ,\varphi `$ ลi nu ลi asupra variabilei $`r`$. รn loc de a caracteriza vectorul $`r`$ prin componentele carteziene $`x,y,z`$ sฤ determinฤm punctul corespunzฤtor $`M`$ de razฤ vectoare $`๐ซ`$ prin coordonatele sferice tridimensionale
$$x=r\mathrm{cos}\varphi \mathrm{sin}\theta ,y=r\mathrm{sin}\varphi \mathrm{sin}\theta ,z=r\mathrm{cos}\theta ,$$
(50)
cu
$$r0,0\theta \pi ,0\varphi 2\pi .$$
Fie $`\mathrm{\Phi }(r,\theta ,\varphi )`$ ลi $`\mathrm{\Phi }^{}(r,\theta ,\varphi )`$ funcลฃiile de undฤ ale unei particule รฎn $`\mathrm{\Sigma }`$ ลi $`\mathrm{\Sigma }^{}`$ รฎn care rotaลฃia infinitezimalฤ este datฤ prin $`\delta \alpha `$ รฎn jurul lui $`z`$
$`\mathrm{\Phi }^{}(r,\theta ,\varphi )`$ $`=`$ $`\mathrm{\Phi }(r,\theta ,\varphi +\delta \alpha ),`$ (51)
$`=`$ $`\mathrm{\Phi }(r,\theta ,\varphi )+\delta \alpha {\displaystyle \frac{\mathrm{\Phi }}{\varphi }},`$
sau
$$\mathrm{\Phi }^{}(r,\theta ,\varphi )=(1+i\widehat{l}_z\delta \alpha )\mathrm{\Phi }(r,\theta ,\varphi ).$$
(52)
Rezultฤ cฤ
$$\frac{\mathrm{\Phi }}{\varphi }=i\widehat{l_z}\mathrm{\Phi },\widehat{l}_z=i\frac{}{\varphi }.$$
(53)
Pentru o rotaลฃie infinitezimalฤ รฎn $`x`$
$`\mathrm{\Phi }^{}(r,\theta ,\varphi )`$ $`=`$ $`\mathrm{\Phi }+\delta \alpha \left({\displaystyle \frac{\mathrm{\Phi }}{\theta }}{\displaystyle \frac{\theta }{\alpha }}+{\displaystyle \frac{\mathrm{\Phi }}{\theta }}{\displaystyle \frac{\varphi }{\alpha }}\right),`$ (54)
$`=`$ $`(1+i\widehat{l}_x\delta \alpha )\mathrm{\Phi }(r,\theta ,\varphi ),`$
รฎnsฤ รฎntr-o astfel de rotaลฃie
$$z^{}=z+y\delta \alpha ;z^{}=z+y\delta \alpha ;x^{}=x$$
(55)
ลi din (50) se obลฃine
$`r\mathrm{cos}(\theta +d\theta )`$ $`=`$ $`r\mathrm{cos}\theta +r\mathrm{sin}\theta \mathrm{sin}\varphi \delta \alpha ,`$
$`r\mathrm{sin}\varphi \mathrm{sin}(\theta +d\theta )`$ $`=`$ $`r\mathrm{sin}\theta \mathrm{sin}\varphi +r\mathrm{sin}\theta \mathrm{sin}\varphi r\mathrm{cos}\theta \delta \alpha ,`$ (56)
adicฤ
$$\mathrm{sin}\theta d\theta =\mathrm{sin}\theta \mathrm{sin}\varphi \delta \alpha \frac{d\theta }{d\alpha }=\mathrm{sin}\varphi ,$$
(57)
ลi
$`\mathrm{cos}\theta \mathrm{sin}\varphi d\theta +\mathrm{sin}\theta \mathrm{cos}\varphi d\varphi `$ $`=`$ $`\mathrm{cos}\theta \delta \alpha ,`$
$`\mathrm{cos}\varphi \mathrm{sin}\theta {\displaystyle \frac{d\varphi }{d\alpha }}`$ $`=`$ $`\mathrm{cos}\theta \mathrm{cos}\theta \mathrm{sin}\varphi {\displaystyle \frac{d\theta }{d\alpha }}.`$ (58)
Substituind (57) รฎn (56) duce la
$$\frac{d\varphi }{d\alpha }=\mathrm{cot}\theta \mathrm{cos}\varphi .$$
(59)
Cu (56) ลi (58) substituite รฎn (51) ลi comparรฎnd pฤrลฃile din dreapta din (51) se obลฃine
$$\widehat{l}_x=i\left(\mathrm{sin}\varphi \frac{}{\theta }+\mathrm{cot}\theta \mathrm{cos}\varphi \frac{}{\varphi }\right).$$
(60)
รn cazul rotaลฃiei รฎn $`y`$, rezultatul este similar
$$\widehat{l}_y=i\left(\mathrm{cos}\varphi \frac{}{\theta }+\mathrm{cot}\theta \mathrm{sin}\varphi \frac{}{\varphi }\right).$$
(61)
Folosind $`\widehat{l}_x,\widehat{l}_y`$ se pot obลฃine deasemenea $`\widehat{l}_\pm ,\widehat{l}^2`$
$`\widehat{l}_\pm `$ $`=`$ $`\mathrm{exp}\left[\pm i\varphi \left(\pm {\displaystyle \frac{}{\theta }}+i\mathrm{cot}\theta {\displaystyle \frac{}{\varphi }}\right)\right],`$
$`\widehat{l}^2`$ $`=`$ $`\widehat{l}_{}\widehat{l}_++\widehat{l}^2+\widehat{l}_z,`$ (62)
$`=`$ $`\left[{\displaystyle \frac{1}{\mathrm{sin}^2\theta }}{\displaystyle \frac{^2}{\varphi ^2}}+{\displaystyle \frac{1}{\mathrm{sin}\theta }}{\displaystyle \frac{}{\theta }}\left(\mathrm{sin}\theta {\displaystyle \frac{}{\theta }}\right)\right].`$
Din (62) se vede cฤ $`\widehat{l}^2`$ este identic pรฎnฤ la o constantฤ cu partea unghiularฤ a operatorului Laplace la o razฤ fixฤ
$$^2f=\frac{1}{r^2}\frac{}{r}\left(r^2\frac{f}{r}\right)+\frac{1}{r^2}\left[\frac{1}{\mathrm{sin}\theta }\frac{}{\theta }\left(\mathrm{sin}\theta \frac{f}{\theta }\right)+\frac{1}{\mathrm{sin}^2\theta }\frac{^2}{\varphi ^2}\right].$$
(63)
### Funcลฃii proprii ale lui $`l_z`$
$`\widehat{l}_z\mathrm{\Phi }_m=m\mathrm{\Phi }=i{\displaystyle \frac{\mathrm{\Phi }_m}{\varphi }},`$
$`\mathrm{\Phi }_m={\displaystyle \frac{1}{\sqrt{2\pi }}}e^{im\varphi }.`$ (64)
### Condiลฃii de hermiticitate ale lui $`\widehat{l}_z`$
$$_0^{2\pi }f^{}\widehat{l}_zg๐\varphi =\left(_0^{2\pi }g^{}\widehat{l}_zf๐\varphi \right)^{}+f^{}g(2\pi )f^{}g(0).$$
(65)
Rezultฤ cฤ $`\widehat{l}_z`$ este hermitic รฎn clasa de funcลฃii pentru care
$$f^{}g(2\pi )=f^{}g(0).$$
(66)
Funcลฃiile proprii $`\mathrm{\Phi }_m`$ ale lui $`\widehat{l}_z`$ aparลฃin clasei integrabile $`^2(0,2\pi )`$ ลi รฎndeplinesc (66), cum se รฎntรฎmplฤ pentru orice funcลฃie care se poate dezvolta รฎn $`\mathrm{\Phi }_m(\varphi )`$
$`F(\varphi )`$ $`=`$ $`{\displaystyle \stackrel{k}{}}a_ke^{ik\varphi },k=0,\pm 1,\pm 2,\mathrm{},`$
$`G(\varphi )`$ $`=`$ $`{\displaystyle \stackrel{k}{}}b_ke^{ik\varphi },k=\pm 1/2,\pm 3/2,\pm 5/2\mathrm{},`$ (67)
cu $`k`$ numai รฎntregi sau numai semiรฎntregi, dar nu pentru combinaลฃii de $`F(\varphi )`$ ลi $`G(\varphi )`$.
Alegerea corectฤ a lui $`m`$ se bazeazฤ รฎn FP comune ale lui $`\widehat{l}_z`$ ลi $`\widehat{l}^2`$.
### Armonice sferice
รn reprezentarea $`\{\stackrel{}{๐ซ}\}`$, funcลฃiile proprii asociate cu autovalorile $`l(l+1)\mathrm{}^2`$, ale lui $`๐ฅ^\mathrm{๐}`$ ลi $`m\mathrm{}`$ ale lui $`l_z`$ sunt soluลฃii ale ecuaลฃiilor diferenลฃiale parลฃiale
$`\left({\displaystyle \frac{^2}{\theta ^2}}+{\displaystyle \frac{1}{\mathrm{tan}\theta }}{\displaystyle \frac{}{\theta }}+{\displaystyle \frac{1}{\mathrm{sin}^2\theta }}{\displaystyle \frac{^2}{\varphi ^2}}\right)\psi (r,\theta ,\varphi )`$ $`=`$ $`l(l+1)\mathrm{}^2\psi (r,\theta ,\varphi ),`$
$`i{\displaystyle \frac{}{\varphi }}\psi (r,\theta ,\varphi )`$ $`=`$ $`m\mathrm{}\psi (r,\theta ,\varphi ).`$ (68)
Considerรฎnd cฤ rezultatele generale obลฃinute sunt aplicabile la cazul momentului cinetic, ลtim cฤ $`l`$ este un รฎntreg sau semiรฎntreg ลi cฤ pentru $`l`$ fixaลฃi $`m`$ poate lua numai valorile
$$l,l+1,\mathrm{},l1,l.$$
รn (68), $`r`$ nu apare รฎn operatorul diferenลฃial, aลa cฤ se poate considera ca un parametru ลi se poate ลฃine cont numai de dependenลฃa รฎn $`\theta ,\varphi `$ a lui $`\psi `$. Se foloseลe notaลฃia $`Y_{lm}(\theta ,\varphi )`$ pentru o astfel de funcลฃie proprie comunฤ a lui $`๐ฅ^\mathrm{๐}`$ ลi $`l_z`$, corespunzฤtoare autovalorilor $`l(l+1)\mathrm{}^2,m\mathrm{}`$ ลi se numeลte armonicฤ sfericฤ .
$`๐ฅ^\mathrm{๐}Y_{lm}(\theta ,\varphi )`$ $`=`$ $`l(l+1)\mathrm{}^2Y_{lm}(\theta ,\varphi ),`$
$`l_zY_{lm}(\theta ,\varphi )`$ $`=`$ $`m\mathrm{}Y_{lm}(\theta ,\varphi ).`$ (69)
Pentru a fi cรฎt mai riguroลi, trebuie sฤ introducem un indice adiลฃional cu scopul de a distinge รฎntre diferite soluลฃii ale lui (69), care corespund aceleaลi perechi de valori $`l,m`$. รntr-adevฤr, dupฤ cum se va vedea mai departe, aceste ecuaลฃii au o soluลฃie unicฤ (pรฎnฤ la un factor constant) pentru fiecare pereche de valori permise ale lui $`l,m`$; aceasta pentru cฤ subindicii $`l,m`$ sunt suficienลฃi รฎn contextul respectiv. Soluลฃiile $`Y_{lm}(\theta ,\varphi )`$ au fost gฤsite prin metoda separฤrii variabilelor รฎn coordonate sferice (vezi capitolul Atomul de hidrogen)
$$\psi _{lm}(r,\theta ,\varphi )=f(r)\psi _{lm}(\theta ,\varphi ),$$
(70)
unde $`f(r)`$ este o funcลฃie de $`r`$, care apare ca o constantฤ de integrare din punctul de vedere al ecuaลฃiilor diferenลฃiale parลฃiale din (68). Faptul cฤ $`f(r)`$ este arbitrarฤ aratฤ cฤ $`๐^\mathrm{๐}`$ ลi $`l_z`$ nu formeazฤ un set complet de observabile<sup>2</sup><sup>2</sup>2Prin definiลฃie, operatorul hermitic A este o observabilฤ dacฤ acest sistem ortogonal de vectori reprezintฤ o bazฤ รฎn spaลฃiul configuraลฃiilor de stฤri. รฎn spaลฃiul $`\epsilon _r`$<sup>3</sup><sup>3</sup>3Fiecare stare cuanticฤ a particulei este caracterizatฤ printr-o stare vectorialฤ aparลฃinรฎnd unui spaลฃiu vectorial abstract $`\epsilon _r`$. al funcลฃiilor de $`\stackrel{}{r}`$ (sau de $`r,\theta ,\varphi `$).
Cu obiectivul de a normaliza $`\psi _{lm}(r,\theta ,\varphi )`$, este convenabil sฤ se normalizeze $`Y_{lm}(\theta ,\varphi )`$ ลi $`f(r)`$ รฎn mod separat:
$`{\displaystyle _0^{2\pi }}๐\varphi {\displaystyle _0^\pi }\mathrm{sin}\theta |\psi _{lm}(\theta ,\varphi )|^2d\theta `$ $`=`$ $`1,`$ (71)
$`{\displaystyle _0^{\mathrm{}}}r^2|f(r)|^2๐r`$ $`=`$ $`1.`$ (72)
### Valorile perechii $`l,m`$
($`\alpha `$): $`l,m`$ trebuie sฤ fie รฎntregi
Folosind $`l_z=\frac{\mathrm{}}{i}\frac{}{\varphi }`$, putem scrie (69) dupฤ cum urmeazฤ
$$\frac{\mathrm{}}{i}\frac{}{\varphi }Y_{lm}(\theta ,\varphi )=m\mathrm{}Y_{lm}(\theta ,\varphi ),$$
(73)
care aratฤ cฤ
$$Y_{lm}(\theta ,\varphi )=F_{lm}(\theta ,\varphi )e^{im\varphi }.$$
(74)
Dacฤ permitem ca $`0\varphi <2\pi `$, atunci putem acoperi tot spaลฃiul pentru cฤ funcลฃia trebuie sฤ fie continuฤ รฎn orice zonฤ , astfel cฤ
$$Y_{lm}(\theta ,\varphi =0)=Y_{lm}(\theta ,\varphi =2\pi ),$$
(75)
ceea ce implicฤ
$$e^{im\pi }=1.$$
(76)
Aลa cum s-a vฤzut, $`m`$ este un รฎntreg sau semiรฎntreg; รฎn cazul aplicaลฃiei la momentul cinetic orbital, $`m`$ trebuie sฤ fie un รฎntreg. ($`e^{2im\pi }`$ ar fi $`1`$ dacฤ $`m`$ ar fi semiรฎntreg).
($`\beta `$): Pentru o valoare datฤ a lui $`l`$, toate armonicele sferice $`Y_{lm}`$ corespunzฤtoare se pot obลฃine prin metode algebrice folosind
$$l_+Y_{ll}(\theta ,\varphi )=0,$$
(77)
care combinatฤ cu ec. (62) pentru $`l_+`$ duce la
$$\left(\frac{d}{d\theta }l\mathrm{cot}\theta \right)F_{ll}(\theta )=0.$$
(78)
Aceastฤ ecuaลฃie poate fi integratฤ imediat dacฤ notฤm relaลฃia
$$\mathrm{cot}\theta d\theta =\frac{d(\mathrm{sin}\theta )}{\mathrm{sin}\theta }.$$
(79)
Soluลฃia sa generalฤ este
$$F_{ll}=c_l(\mathrm{sin}\theta )^l,$$
(80)
unde $`c_l`$ este o constantฤ de normalizare.
Rezultฤ cฤ pentru orice valoare pozitivฤ sau zero a lui $`l`$, existฤ o funcลฃie $`Y_{ll}(\theta ,\varphi )`$ care este (cu un factor constant asociat)
$$Y^{ll}(\theta ,\varphi )=c_l(\mathrm{sin}\theta )^le^{il\varphi }.$$
(81)
Folosind acลฃiunea lui $`l_{}`$ รฎn mod repetat se poate construi tot setul de funcลฃii $`Y_{ll1}(\theta ,\varphi ),\mathrm{},Y_{l0}(\theta ,\varphi ),`$ $`\mathrm{},Y_{ll}(\theta ,\varphi )`$. รn continuare, vedem modul รฎn care se corespund aceste funcลฃii cu perechea de autovalori $`l(l+1)\mathrm{},m\mathrm{}`$ (unde $`l`$ este un รฎntreg pozitiv arbitrar sau zero ลi $`m`$ este alt รฎntreg astfel cฤ $`lml`$ ). Pe baza lui (78) ajungem la concluzia cฤ orice funcลฃie proprie $`Y_{lm}(\theta ,\varphi )`$, poate fi obลฃinutฤ รฎn mod neambiguu din $`Y_{ll}`$.
### Proprietฤลฃi ale armonicelor sferice
$`\alpha `$ Relaลฃii de recurenลฃฤ
Urmรฎnd rezultatele generale avem
$$l_\pm Y_{lm}(\theta ,\varphi )=\mathrm{}\sqrt{l(l+1)m(m\pm 1)}Y_{lm\pm 1}(\theta ,\varphi ).$$
(82)
Folosind expresia (62) pentru $`l_\pm `$ ลi faptul cฤ $`Y_{lm}(\theta ,\varphi )`$ este produsul รฎntre o funcลฃie dependentฤ numai de $`\theta `$ ลi $`e^{\pm i\varphi }`$, obลฃinem
$$e^{\pm i\varphi }\left(\frac{}{\theta }m\mathrm{cot}\theta \right)Y_{lm}(\theta ,\varphi )=\sqrt{l(l+1)m(m\pm 1)}Y_{lm\pm 1}(\theta ,\varphi )$$
(83)
$`\beta `$ Ortonormare ลi relaลฃii de completitudine
Ecuaลฃia (68) determinฤ numai armonicele sferice pรฎnฤ la un factor constant. Acum vom alege acest factor astfel ca sฤ avem ortonormalizarea funcลฃiilor $`Y_{lm}(\theta ,\varphi )`$ (ca funcลฃii de variabilele unghiulare $`\theta ,\varphi `$)
$$_0^{2\pi }๐\varphi _0^\pi \mathrm{sin}\theta d\theta Y_{lm}^{}(\theta ,\varphi )Y_{lm}(\theta ,\varphi )=\delta _{l^{}l}\delta _{m^{}m}.$$
(84)
รn plus, orice funcลฃie continuฤ de $`\theta ,\varphi `$ pot fi expresate cu ajutorul armonicelor sferice, adicฤ
$$f(\theta ,\varphi )=\underset{l=0}{\overset{\mathrm{}}{}}\underset{m=l}{\overset{l}{}}c_{lm}Y_{lm}(\theta ,\varphi ),$$
(85)
unde
$$c_{lm}=_0^{2\pi }๐\varphi _0^\pi \mathrm{sin}\theta d\theta Y_{lm}^{}(\theta ,\varphi )f(\theta ,\varphi ).$$
(86)
Rezultatele (85), (86) sunt consecinลฃa definirii armonicele sferice ca bazฤ ortonormalฤ completฤ รฎn spaลฃiul $`\epsilon _\mathrm{\Omega }`$ a funcลฃiilor de $`\theta ,\varphi `$. Relaลฃia de completitudine este
$`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=l}{\overset{l}{}}}Y_{lm}(\theta ,\varphi )Y_{lm}^{}(\theta ^{},\varphi )`$ $`=`$ $`\delta (\mathrm{cos}\theta \mathrm{cos}\theta ^{})\delta (\varphi ,\varphi ),`$ (87)
$`=`$ $`{\displaystyle \frac{1}{\mathrm{sin}\theta }}\delta (\theta \theta ^{})\delta (\varphi ,\varphi ).`$
โFuncลฃiaโ $`\delta (\mathrm{cos}\theta \mathrm{cos}\theta ^{})`$ apare pentru cฤ integrala peste variabila $`\theta `$ se efectueazฤ folosind elementul diferenลฃial $`\mathrm{sin}\theta d\theta =d(\mathrm{cos}\theta )`$.
### Operatorul de paritate $`๐ซ`$ รฎn cazul armonicelor sferice
Comportamentul lui $`๐ซ`$ รฎn trei dimensiuni este destul de asemฤnฤtor celui รฎntr-o singurฤ dimensiune, respectiv, cรฎnd se aplicฤ unei funcลฃii $`\psi (x,y,z)`$ de coordonate carteziene รฎi modificฤ numai semnul fiecฤreia dintre coordonate
$$๐ซ\psi (x,y,z)=\psi (x,y,z).$$
(88)
$`๐ซ`$ are proprietฤลฃile unui operator hermitic ลi รฎn plus este un operator unitar precum ลi proiector deoarece $`๐ซ^2`$ este un operator identitate
$`\psi (๐ซ)|๐ซ|\psi (๐ซ)=\psi (๐ซ)|\psi (๐ซ)=\psi (๐ซ)|\psi (๐ซ),`$
$`๐ซ^2|๐ซ=๐ซ(๐ซ|๐ซ=๐ซ|๐ซ=|๐ซ`$ (89)
ลi deci
$$๐ซ^2=\widehat{1},$$
(90)
pentru care valorile proprii sunt $`P=\pm 1`$. FP-urile se numesc pare dacฤ $`P=1`$ ลi impare dacฤ $`P=1`$. รn mecanica cuanticฤ nerelativistฤ , operatorul $`\widehat{H}`$ pentru un sistem conservativ este invariant la transformฤri unitare discrete
$$๐ซ\widehat{H}๐ซ=๐ซ^1\widehat{H}๐ซ=\widehat{H}.$$
(91)
Astfel $`\widehat{H}`$ comutฤ cu $`๐ซ`$ ลi deci paritatea stฤrii este constantฤ de miลcare. Deasemenea $`๐ซ`$ comutฤ cu operatorii $`\widehat{l}`$ ลi $`\widehat{l}_\pm `$
$$[๐ซ,\widehat{l}_i]=0,[๐ซ,\widehat{l}_\pm ]=0.$$
(92)
Dacฤ $`\psi `$ este FP a tripletei $`๐ซ,\widehat{l}`$ ลi $`\widehat{l}_z`$, din (92) rezultฤ cฤ paritฤลฃile stฤrilor care diferฤ numai รฎn $`\widehat{l}_z`$ coincid. รn felul acesta se identificฤ paritatea unei particule de moment unghiular $`\widehat{l}`$.
รn coordonate sferice, pentru acest operator vom considera urmฤtoarea substituลฃie
$$rr,\theta \pi \theta \varphi \pi +\varphi .$$
(93)
Prin urmare, dacฤ folosim o bazฤ standard รฎn spaลฃiul funcลฃiilor de undฤ a unei particule fฤrฤ โrotaลฃie proprieโ, partea radialฤ a funcลฃiilor $`\psi _{klm}(\stackrel{}{r})`$ din bazฤ nu este modificatฤ de cฤtre operatorul paritate. Doar armonicele sferice se schimbฤ . Transformฤrile (93) se traduc trigonometric รฎn:
$$\mathrm{sin}(\pi \theta )\mathrm{sin}\theta ,\mathrm{cos}(\pi \theta )\mathrm{cos}\theta e^{im(\pi +\varphi }(1)^me^{im\varphi }$$
(94)
ลi รฎn aceste condiลฃii, funcลฃia $`Y_{ll}(\theta ,\varphi )`$ se transformฤ รฎn
$$Y_{ll}(\varphi \theta ,\pi +\varphi )=(1)^lY_{ll}(\theta ,\varphi ).$$
(95)
Din (95) rezultฤ cฤ paritatea lui $`Y_{ll}`$ este $`(1)^l`$. Pe de altฤ parte, $`l_{}`$ (ca ลi $`l_+`$ este invariant la transformฤrile:
$$\frac{}{(\pi \theta )}\frac{}{\theta },\frac{}{(\pi +\varphi )}\frac{}{\varphi }e^{i(\pi +\varphi )}e^{i\varphi }\mathrm{cot}(\pi \theta )\mathrm{cot}\theta .$$
(96)
Cu alte cuvinte $`l_\pm `$ sunt pari. Prin urmare, deducem cฤ paritatea oricฤrei armonice sferice este $`(1)^l`$, adicฤ este invariantฤ la schimbฤri azimutale:
$$Y_{lm}(\varphi \theta ,\pi +\varphi )=(1)^lY_{lm}(\theta ,\varphi ).$$
(97)
Aลadar, armonicele sferice sunt funcลฃii de paritate bine definitฤ , independentฤ de $`m`$, parฤ dacฤ $`l`$ este par ลi imparฤ dacฤ $`l`$ este impar.
## Operatorul de spin
Unele particule, posedฤ nu numai moment cinetic รฎn raport cu axe exterioare ci ลi un moment propriu, care se cunoaลte ca spin ลi pe care รฎl vom nota cu $`\widehat{S}`$. Acest operator nu este relaลฃionat cu rotaลฃii normale faลฃฤ de axe โrealeโ รฎn spaลฃiu, dar respectฤ relaลฃii de comutare de acelaลi tip ca ale momentului cinetic orbital, respectiv
$$[\widehat{S}_i,\widehat{S}_j]=i\epsilon _{ijk}\widehat{S}_k,$$
(98)
odatฤ cu urmฤtoarele proprietฤลฃi:
* Pentru spin, se satisfac toate formulele de la (23) pรฎnฤ la (48) ale momentului cinetic.
* Spectrul proiecลฃiilor de spin, este o secvenลฃฤ de numere รฎntregi sau semiรฎntregi care diferฤ printr-o unitate.
* Valorile proprii ale lui $`\widehat{S}^2`$ sunt
$$\widehat{S}^2\psi _s=S(S+1)\psi _s.$$
(99)
* Pentru un $`S`$ dat, componentele $`S_z`$ pot sฤ ia numai $`2S+1`$ valori, de la $`S`$ la $`+S`$.
* FP-urile particulelor cu spin, pe lรฎngฤ dependenลฃa de $`\stackrel{}{r}`$ ลi/sau $`\stackrel{}{p}`$, depind de o variabilฤ discretฤ (proprie spinului) $`\sigma `$, care denotฤ proiecลฃia spinului pe axa $`z`$.
* FP-urile $`\psi (\stackrel{}{r},\sigma )`$ ale unei particule cu spin se pot dezvolta รฎn FP-uri cu proiecลฃii date ale spinului $`S_z`$, adicฤ
$$\psi (\stackrel{}{r},\sigma )=\underset{\sigma =S}{\overset{S}{}}\psi _\sigma (\stackrel{}{r})\chi (\sigma ),$$
(100)
unde $`\psi _\sigma (\stackrel{}{r})`$ este partea orbitalฤ ลi $`\chi (\sigma )`$ este partea spinorialฤ .
* Funcลฃiile de spin (spinorii) $`\chi (\sigma _i)`$ sunt ortogonale pentru orice pereche $`\sigma _i\sigma _k`$. Funcลฃiile $`\psi _\sigma (\stackrel{}{r})\chi (\sigma )`$ din suma (100) sunt componentele unei FP a unei particule cu spin.
* Funcลฃia $`\psi _\sigma (\stackrel{}{r})`$ se numeลte parte orbitalฤ a spinorului, sau mai scurt orbital.
* Normalizarea spinorilor se face รฎn felul urmฤtor
$$\underset{\sigma =S}{\overset{S}{}}\psi _\sigma (\stackrel{}{r})=1.$$
(101)
Relaลฃiile de comutare permit sฤ se stabileascฤ forma concretฤ a operatorilor (matricelor) de spin care acลฃioneazฤ รฎn spaลฃiul FP-urilor operatorului proiecลฃiei spinului.
Multe particule โelementareโ, cum ar fi electronul, neutronul, protonul, etc. au spin $`1/2`$ ลi de aceea proiecลฃia spinului lor ia numai douฤ valori, respectiv $`S_z=\pm 1/2`$ (รฎn unitฤลฃi $`\mathrm{}`$). Fac parte din clasa fermionilor, datoritฤ statisticii pe care o prezintฤ cรฎnd formeazฤ sisteme de multe corpuri.
Pe de altฤ parte, matricele $`S_x,S_y,S_z`$ รฎn spaลฃiul FP-urilor lui $`\widehat{S}^2,\widehat{S}_z`$ sunt
$`S_x={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),`$ $`S_y={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),`$ (106)
$`S_z={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),`$ $`S^2={\displaystyle \frac{3}{4}}\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).`$ (111)
### Definiลฃia matricelor Pauli
Matricele
$$\sigma _i=2S_i$$
(112)
se numesc matricele (lui) Pauli. Sunt matrici hermitice ลi au aceeaลi ecuaลฃie caracteristicฤ
$$\lambda ^21=0,$$
(113)
ลi prin urmare, autovalorile lui $`\sigma _x,\sigma _y`$ ลi $`\sigma _z`$ sunt
$$\lambda =\pm 1.$$
(114)
Algebra acestor matrici este urmฤtoarea:
$$\sigma _i^2=\widehat{I},\sigma _k\sigma _j=\sigma _j\sigma _k=i\sigma _z,\sigma _j\sigma _k=i\underset{l}{}\epsilon _{jkl}\sigma _l.+\delta _{jk}I.$$
(115)
รn cazul รฎn care sistemul cu spin are simetrie sfericฤ
$$\psi _1(r,+\frac{1}{2}),\psi _1(r,\frac{1}{2}),$$
(116)
sunt soluลฃii diferite datoritฤ proiecลฃiilor $`S_z`$ diferite. Valoarea probabilitฤลฃii uneia sau alteia dintre proiecลฃii este determinatฤ de modulul pฤtrat $`\psi _1^2`$ sau $`\psi _2^2`$ รฎn aลa fel cฤ
$$\psi _1^2+\psi _2^2=1.$$
(117)
Cum FP ale lui $`S_z`$ are douฤ componente, atunci
$$\chi _1=\left(\begin{array}{c}1\\ 0\end{array}\right),\chi _2=\left(\begin{array}{c}0\\ 1\end{array}\right),$$
(118)
astfel cฤ FP a unei particule de spin $`1/2`$ se poate scrie ca o matrice de o columnฤ
$$\psi =\psi _1\chi _1+\psi _2\chi _2=\left(\begin{array}{c}\psi _1\\ \psi _2\end{array}\right).$$
(119)
รn continuare, orbitalii vor fi substituiลฃi prin numere datoritฤ faptului cฤ ne intereseazฤ numai partea de spin.
## Transformฤrile la rotaลฃii ale spinorilor
Fie $`\psi `$ funcลฃia de undฤ a unui sistem cu spin รฎn $`\mathrm{\Sigma }`$. Sฤ determinฤm probabilitatea proiecลฃiei spinului รฎntr-o direcลฃie arbitrarฤ รฎn spaลฃiul tridimensional (3D) care se poate alege ca axฤ $`z^{}`$ a lui $`\mathrm{\Sigma }^{}`$. Cum am vฤzut deja pentru cazul momentului cinetic, existฤ douฤ metode de abordare ลi soluลฃionare a acestei probleme:
* $`\psi `$ nu se schimbฤ cรฎnd $`\mathrm{\Sigma }\mathrm{\Sigma }^{}`$ ลi operatorul $`\widehat{\mathrm{\Lambda }}`$ se transformฤ ca un vector. Trebuie sฤ gฤsim FP-urile proiecลฃiilor $`S_z^{}`$ ลi sฤ dezvoltฤm $`\psi `$ รฎn aceste FP-uri. Pฤtratele modulelor coeficienลฃilor dau rezultatul
$`\widehat{S}_x^{}=\widehat{S}_x\mathrm{cos}\varphi +\widehat{S}_y\mathrm{sin}\varphi `$ $`=`$ $`e^{il\varphi }\widehat{S}_xe^{il\varphi },`$
$`\widehat{S}_y^{}=\widehat{S}_x\mathrm{sin}\varphi +\widehat{S}_y\mathrm{cos}\varphi `$ $`=`$ $`e^{il\varphi }\widehat{S}_ye^{il\varphi },`$
$`\widehat{S}_z^{}=\widehat{S}_z=e^{il\varphi }\widehat{S}_z,`$ (120)
cu rotaลฃii infinitezimale ลi din relaลฃiile de comutare pentru spin se poate gฤsi
$$\widehat{l}=\widehat{S}_z,$$
(121)
unde $`\widehat{l}`$ este generatorul infinitezimal.
* A doua reprezentare este:
$`\widehat{S}`$ nu se schimbฤ cรฎnd $`\mathrm{\Sigma }\mathrm{\Sigma }^{}`$ ลi componentele lui $`\psi `$ se schimbฤ . Transformarea la aceastฤ reprezentare se face cu o transformare unitarฤ de forma
$`\widehat{V}^{}\widehat{S}^{}\widehat{V}`$ $`=`$ $`\widehat{\mathrm{\Lambda }},`$
$`\left(\begin{array}{c}\psi _1^{}\\ \psi _2^{}\end{array}\right)`$ $`=`$ $`\widehat{V}^{}\left(\begin{array}{c}\psi _1\\ \psi _2\end{array}\right).`$ (126)
Pe baza lui (111) ลi (113) rezultฤ cฤ
$`\widehat{V}^{}e^{i\widehat{S}_z\varphi }\widehat{S}e^{i\widehat{S}_z\varphi }\widehat{V}`$ $`=`$ $`\widehat{S},`$
$`\widehat{V}^{}`$ $`=`$ $`e^{i\widehat{S}_z\varphi },`$ (127)
ลi din (114) se obลฃine
$$\left(\begin{array}{c}\psi _1^{}\\ \psi _2^{}\end{array}\right)=e^{i\widehat{S}_z\varphi }\left(\begin{array}{c}\psi _1\\ \psi _2\end{array}\right).$$
(128)
Folosind forma concretฤ a lui $`\widehat{S}_z`$ ลi proprietฤลฃile matricelor Pauli se obลฃine forma concretฤ $`\widehat{V}_z^{}`$, astfel cฤ
$$\widehat{V}_z^{}(\varphi )=\left(\begin{array}{cc}e^{\frac{i}{2}\varphi }& 0\\ 0& e^{\frac{i}{2}\varphi }\end{array}\right).$$
(129)
## Un rezultat al lui Euler
Se poate ajunge la orice sistem de referinลฃฤ $`\mathrm{\Sigma }^{}`$ de orientare arbitrarฤ faลฃฤ de $`\mathrm{\Sigma }`$ prin numai trei rotaลฃii, prima de unghi $`\varphi `$ รฎn jurul axei $`z`$, urmฤtoarea rotaลฃie de unghi $`\theta `$ รฎn jurul noii axe de coordonate $`x^{}`$ ลi ultima de unghi $`\psi _a`$ รฎn jurul lui $`z^{}`$. Acest rezultat important aparลฃine lui Euler.
Parametrii $`(\phi ,\theta ,\psi _a)`$ se numesc unghiurile (lui) Euler
$$\widehat{V}^{}(\phi ,\theta ,\psi _a)=\widehat{V}_z^{}^{}(\psi _a)\widehat{V}_x^{}^{}(\theta )\widehat{V}_z^{}(\phi ).$$
(130)
Matricele $`\widehat{V}_z^{}`$ sunt de forma (116), รฎn timp ce $`\widehat{V}_x^{}`$ este de forma
$$\widehat{V}_x^{}(\phi )=\left(\begin{array}{cc}\mathrm{cos}\frac{\theta }{2}& i\mathrm{sin}\frac{\theta }{2}\\ i\mathrm{sin}\frac{\theta }{2}& \mathrm{cos}\frac{\theta }{2}\end{array}\right),$$
(131)
astfel cฤ
$$\widehat{V}^{}(\phi ,\theta ,\psi _a)=\left(\begin{array}{cc}e^{i\frac{\phi +\psi _a}{2}}\mathrm{cos}\frac{\theta }{2}& ie^{i\frac{\psi _a\phi }{2}}\mathrm{sin}\frac{\theta }{2}\\ ie^{i\frac{\phi \psi _a}{2}}\mathrm{sin}\frac{\theta }{2}& e^{i\frac{\phi +\psi _a}{2}}\mathrm{cos}\frac{\theta }{2}\end{array}\right).$$
(132)
Rezultฤ deci cฤ prin rotaลฃia lui $`\mathrm{\Sigma }`$, componentele funcลฃiei spinoriale se transformฤ dupฤ cum urmeazฤ
$`\psi _1^{}`$ $`=`$ $`\psi _1e^{i\frac{\phi +\psi _a}{2}}\mathrm{cos}{\displaystyle \frac{\theta }{2}}+i\psi _2e^{i\frac{\psi _a\phi }{2}}\mathrm{sin}{\displaystyle \frac{\theta }{2}},`$
$`\psi _2^{}`$ $`=`$ $`i\psi _1e^{i\frac{\phi \psi _a}{2}}\mathrm{sin}{\displaystyle \frac{\theta }{2}}+\psi _2e^{i\frac{\phi +\psi _a}{2}}\mathrm{cos}{\displaystyle \frac{\theta }{2}}.`$ (133)
Din (120) se poate vedea cฤ unei rotaลฃii รฎn $`E_3`$ รฎi corespunde o transformare linearฤ รฎn $`E_2`$, spaลฃiul euclidean bidimensional, relaลฃionatฤ cu cele douฤ componente ale funcลฃiei spinoriale. Rotaลฃia รฎn $`E_3`$ nu implicฤ o rotaลฃie รฎn $`E_2`$, ceea ce รฎnseamnฤ
$$\mathrm{\Phi }^{}|\psi ^{}=\mathrm{\Phi }|\psi =\mathrm{\Phi }_1^{}\psi _1+\mathrm{\Phi }_2^{}\psi _2.$$
(134)
Din (119) se obลฃine cฤ (121) nu se satisface, totuลi existฤ o invarianลฃฤ รฎn transformฤrile (119) รฎn spaลฃiul $`E_2`$ al funcลฃiilor spinoriale,
$$\{\mathrm{\Phi }|\psi \}=\psi _1\mathrm{\Phi }_2\psi _2\mathrm{\Phi }_1.$$
(135)
Transformฤrile lineare care menลฃin invariante astfel de forme bilineare se numesc binare.
O mฤrime fizicฤ cu douฤ componente pentru care o rotaลฃie a sistemului de coordonate este o transformare binarฤ se numeลte spin de ordinul รฎntรฎi sau pe scurt spin.
### Spinorii unui sistem de doi fermioni
Funcลฃiile proprii ale lui $`{}_{i}{}^{}\widehat{s}_{i}^{2}\widehat{s}_z`$, cu $`i=1,2`$ au urmฤtoarea formฤ
$$i|+=\left(\begin{array}{c}1\\ 0\end{array}\right)_i,i|=\left(\begin{array}{c}0\\ 1\end{array}\right)_i.$$
(136)
Un operator foarte folosit รฎntr-un sistem de doi fermioni este spinul total
$$\widehat{S}=_1\widehat{S}+_2\widehat{S}$$
(137)
Spinorii lui $`\widehat{s}^2\widehat{s}_z`$ sunt ket-uri $`|\widehat{S},\sigma `$, care sunt combinaลฃii lineare ale spinorilor $`{}_{i}{}^{}\widehat{s}_{i}^{2}\widehat{s}_z`$
$`|++=(\begin{array}{c}1\\ 0\end{array}\left)_1\right(\begin{array}{c}1\\ 0\end{array})_1,`$ $`|+=(\begin{array}{c}1\\ 0\end{array}\left)_1\right(\begin{array}{c}0\\ 1\end{array})_2,`$ (146)
$`|+=(\begin{array}{c}0\\ 1\end{array}\left)_2\right(\begin{array}{c}1\\ 0\end{array})_1,`$ $`|=(\begin{array}{c}0\\ 1\end{array}\left)_2\right(\begin{array}{c}0\\ 1\end{array})_2.`$ (155)
Funcลฃiile spinoriale din (125) se considerฤ ortonormalizate. รn $`E_n`$ ket-ul $`|++`$ este de $`S_z=1`$ ลi รฎn acelaลi timp este funcลฃie proprie a operatorului
$$\widehat{S}=_1\widehat{s}^2+2(_1\widehat{s})(_2\widehat{s})+_2\widehat{s}^2.$$
(156)
Dupฤ cum se poate vedea din
$`\widehat{S}^2`$ $`=`$ $`|++=\frac{3}{2}|+++2(_1\widehat{s}_x_2\widehat{s}_x+_1\widehat{s}_y_2\widehat{s}_y+_1\widehat{s}_z_2\widehat{s}_z)|++,`$ (157)
$`\widehat{S}^2`$ $`=`$ $`|++=2|++=1(1+1)|++.`$ (158)
Dacฤ se introduce operatorul
$$\widehat{S}_{}=_1\widehat{s}_{}+_2\widehat{s}_{},$$
(159)
se obลฃine cฤ
$$[\widehat{S}_{},\widehat{S}^2]=0.$$
(160)
Atunci $`(\widehat{S}_{})^k|1,1`$ se poate scrie รฎn funcลฃie de FP-urile operatorului $`\widehat{S}^2`$, respectiv
$$\widehat{S}_{}|1,1=\widehat{S}_{}|++=\sqrt{2}|++\sqrt{2}|+.$$
(161)
Rezultฤ cฤ $`S_z=0`$ รฎn starea $`\widehat{S}_{}|1,1`$. Pe de altฤ parte, din condiลฃia de normalizare avem
$`|1,0=\frac{1}{\sqrt{2}}(|++|+)`$ (162)
$`\widehat{S}_{}|1,0=|+|=\alpha |1,1.`$ (163)
Din condiลฃia de normalizare
$$|1,1=|,.$$
(164)
Existฤ รฎncฤ o singurฤ combinaลฃie linear independentฤ de funcลฃii de tip (125) diferitฤ de $`|1,1,|1,0`$ y $`|1,1`$, respectiv
$`\psi _4=\frac{1}{\sqrt{2}}(|+|+),`$ (165)
$`\widehat{S}_z\psi _4=0,\widehat{S}^2\psi _4.`$ (166)
Prin urmare
$$\psi _4=|0,0.$$
(167)
$`\psi _4`$ descrie starea unui sistem de doi fermioni cu spinul total egal zero. Acest tip de stare se numeลte singlet. Pe de altฤ parte, starea a doi fermioni de spin total egal unu se poate numi triplet avรฎnd un grad de degenerare $`g=3`$.
## Moment unghiular total
Momentul unghiular total este un operator care se introduce ca suma momentului unghiular orbital ลi de spin, respectiv
$$\widehat{J}=\widehat{l}+\widehat{S},$$
(168)
unde $`\widehat{l}`$ ลi $`\widehat{S}`$, aลa cum am vฤzut, acลฃioneazฤ รฎn spaลฃii diferite, dar pฤtratele lui $`\widehat{l}`$ ลi $`\widehat{S}`$ comutฤ cu $`\widehat{J}`$, adicฤ
$$[\widehat{J}_i,\widehat{J}_j]=i\epsilon _{ijk}\widehat{J}_k,[\widehat{J}_i,\widehat{l}^2]=0,[\widehat{J}_i,\widehat{S}^2]=0,$$
(169)
Din (139) rezultฤ cฤ $`\widehat{l}^2`$ ลi $`\widehat{S}^2`$ au un sistem comun de FP-uri cu $`\widehat{J}^2`$ ลi $`\widehat{J}_z`$.
Sฤ determinฤm spectrul proiecลฃiilor lui $`\widehat{J}_z`$ pentru un fermion. Starea de proiecลฃie de $`\widehat{J}_z`$ maxim se poate scrie
$`\overline{\psi }`$ $`=`$ $`\psi _{ll}\left(\begin{array}{c}1\\ 0\end{array}\right)=|l,l,+`$ (172)
$`\widehat{ศท}_z\psi `$ $`=`$ $`(l+\frac{1}{2})\overline{\psi },j=l+\frac{1}{2}.`$ (173)
Introducem operatorul $`\widehat{J}_{}`$ definit prin
$$\widehat{J}_{}=\widehat{l}_{}+\widehat{S}_{}=\widehat{l}_{}+\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right).$$
(174)
Pe baza normalizฤrii $`\alpha =\sqrt{(J+M)(JM+1)}`$ se obลฃine
$$\widehat{J}_{}\psi _{ll}\left(\begin{array}{c}1\\ 0\end{array}\right)=\sqrt{2l}|l,l1,++|l,l1,,$$
(175)
astfel cฤ valoarea proiecลฃiei lui $`\widehat{j}_{}`$ รฎn $`\widehat{j}_{}\overline{\psi }`$ va fi
$$\widehat{ศท}_z=(l1)+\frac{1}{2}=l\frac{1}{2}.$$
(176)
Rezultฤ cฤ $`\widehat{ศท}_{}`$ micลoreazฤ cu o unitate acลฃiunea lui $`\widehat{J}_z`$.
รn cazul general avem
$$\widehat{ศท}_{}^k=\widehat{l}_{}^k+k\widehat{l}_{}^{k1}\widehat{S}_{}.$$
(177)
Se observฤ cฤ (145) se obลฃine din dezvoltarea binomialฤ considerรฎnd รฎn plus cฤ $`\widehat{s}_{}^2`$ ลi toate puterile superioare ale lui $`\widehat{s}`$ sunt zero.
$$\widehat{ศท}_{}^k|l,l,+=\widehat{l}_{}^k|l,l,++k\widehat{l}_{}^{k1}|l,l,.$$
(178)
ลtim cฤ
$$(\widehat{l}_{})^k\psi _{l,l}=\sqrt{\frac{k!(2l)!}{(2lk)!}}\psi _{l,lk}$$
ลi folosind-o se obลฃine
$$\widehat{ศท}_{}^k|l,l,+=\sqrt{\frac{k!(2l)!}{(2lk)!}}|l,lk,++\sqrt{\frac{(k+1)!(2l)!}{(2lk+1)!}}k|l,lk+1,.$$
(179)
Notฤm acum $`m=lk`$
$$\widehat{ศท}_{}^{lm}|l,l,+=\sqrt{\frac{(lm)!(2l)!}{(l+m)!}}|l,m,++\sqrt{\frac{(lm1)!(2l)!}{(2l+m+1)!}}(lm)|l,m+1,.$$
(180)
Valorile proprii ale proiecลฃiei momentului unghiular total sunt date de secvenลฃa de numere care diferฤ printr-o unitate de la $`j=l+\frac{1}{2}`$ pรฎnฤ la $`j=l\frac{1}{2}`$. Toate aceste stฤri aparลฃin aceleaลi funcลฃii proprii a lui $`\widehat{J}`$ ca ลi $`|l,l,+`$ pentru cฤ $`[\widehat{J}_{},\widehat{J}^2]=0`$
$`\widehat{J}^2|l,l,+`$ $`=`$ $`(\widehat{l}^2+2\widehat{l}\widehat{S}+\widehat{S}^2)|l,l,+,`$ (181)
$`=`$ $`[l(l+1)+2l\frac{1}{2}+\frac{3}{4}]|l,l,+`$
unde $`j(j+1)=(l+\frac{1}{2})(l+\frac{3}{2})`$.
รn partea dreaptฤ a lui (149) o contribuลฃie diferitฤ de zero dฤ numai $`j=\widehat{l}_z\widehat{S}_z`$. Atunci FP-urile obลฃinute corespund perechii $`j=l+\frac{1}{2}`$, $`m_j=m+\frac{1}{2}`$ ลi sunt de forma
$$|l+\frac{1}{2},m+\frac{1}{2}=\sqrt{\frac{l+m+1}{2l+1}}|l,m,++\sqrt{\frac{lm}{2l+1}}|l,m+1,.$$
(182)
Numฤrul total de stฤri linear independente este
$$N=(2l+1)(2s+1)=4l+2,$$
(183)
din care รฎn (150) s-au construit (2j+1)=2l+3. Restul de $`2l1`$ funcลฃii proprii se pot obลฃine din condiลฃia de ortonormalizare:
$$|l\frac{1}{2},m\frac{1}{2}=\sqrt{\frac{lm}{2l+1}}|l,m,+\sqrt{\frac{l+m+1}{2l+1}}|l,m+1,.$$
(184)
Dacฤ douฤ subsisteme sunt รฎn interacลฃiune รฎn aลa fel รฎncรฎt fiecare moment unghiular $`\widehat{j}_i`$ se conservฤ , atunci FP-urile operatorului moment unghiular total
$$\widehat{J}=\widehat{ศท}_1+\widehat{ศท}_2,$$
(185)
se pot obลฃine printr-o procedurฤ asemฤnฤtoare celei anterioare. Pentru valori proprii fixe ale lui $`\widehat{ศท}_1`$ ลi $`\widehat{ศท}_2`$ existฤ $`(2j_1+1)(2j_2+1)`$ FP-uri ortonormalizate ale proiecลฃiei momentului unghiular total $`\widehat{J}_z`$, iar cea care corespunde valorii maxime a proiecลฃiei $`\widehat{J}_z`$, adicฤ $`M_J=j_1+j_2`$, se poate construi รฎn mod unic ลi prin urmare $`J=j_1+j_2`$ este valoarea maximฤ a momentului unghiular total al sistemului. Aplicรฎnd operatorul $`\widehat{J}=\widehat{ศท}_1+\widehat{ศท}_2`$ รฎn mod repetat funcลฃiei
$$|j_1+j_2,j_1+j_2,j_1+j_2=|j_1,j_1|j_2,j_2,$$
(186)
se pot obลฃine toate cele $`2(j_1+j_2)+1`$ FP ale lui $`\widehat{J}=j_1+j_2`$ cu diferiลฃi $`M`$:
$$(j_1+j_2)M(j_1+j_2).$$
De exemplu, FP pentru $`M=j_1+j_21`$ este:
$$|j_1+j_2,j_1+j_21,j_1,j_2=\sqrt{\frac{j_1}{j_1+j_2}}|j_1,j_11,j_2,j_2+\sqrt{\frac{j_2}{j_1+j_2}}|j_1,j_1,j_2,j_21.$$
(187)
Aplicรฎnd รฎn continuare de mai multe ori operatorul $`\widehat{J}_{}`$ se pot obลฃine cele $`2(j_1+j_21)1`$ funcลฃii ale lui $`J=j_1+j_21`$.
Se poate demonstra cฤ
$$|j_1j_2|Jj_1+j_2$$
astfel cฤ
$$\underset{\mathrm{min}J}{\overset{\mathrm{max}J}{}}(2J+1)=(2J_1+1)(2J_2+1)$$
(188)
ลi deci
$$|J,M,j_1,j_2=\underset{m_1+m_2=M}{}(j_1m_1j_2m_2|JM)|j_1,m_1,j_2,m_2,$$
(189)
unde coeficienลฃii $`(j_1m_1j_2m_2|JM)`$ determinฤ contribuลฃia diferitelor funcลฃii $`|j_1,m_1,j_2,m_2`$ รฎn funcลฃiile proprii ale lui $`\widehat{J^2}`$, $`\widehat{J_z}`$ de valori proprii $`J(J+1)`$, $`M`$ ลi sunt numiลฃi coeficienลฃii Clebsch-Gordan.
Referinลฃe:
1. H.A. Buchdahl, โRemark concerning the eigenvalues of orbital angular momentumโ,
Am. J. Phys. 30, 829-831 (1962)
3N. Notฤ: 1. Operatorul corespunzฤtor vectorului Runge-Lenz din problema Kepler clasicฤ se scrie
$$\widehat{\stackrel{}{A}}=\frac{\widehat{๐ซ}}{r}+\frac{1}{2}\left[(\widehat{l}\times \widehat{p})(\widehat{p}\times \widehat{l})\right].$$
unde s-au folosit unitฤลฃi atomice ลi s-a considerat $`Z=1`$ (atomul de hidrogen). Acest operator comutฤ cu Hamiltonianul atomului de hidrogen $`\widehat{H}=\frac{\widehat{p^2}}{2}\frac{1}{r}`$, adicฤ este integralฤ de miลcare cuanticฤ . Componentele sale au comutatori de tipul $`[A_i,A_j]=2iฯต_{ijk}l_kH`$, iar comutatorii componentelor Runge-Lenz cu componentele momentului cinetic sunt de tipul $`[l_i,A_j]=iฯต_{ijk}A_k`$, adicฤ respectฤ condiลฃiile (23). De demonstrat toate aceste relaลฃii poate fi un exerciลฃiu util.
## 3P. Probleme
### Problema 3.1
Sฤ se arate cฤ orice operator de translaลฃie, pentru care $`\psi (y+a)=T_a\psi (y)`$, se poate scrie ca un operator exponenลฃial ลi sฤ se aplice acest rezultat pentru $`y=\stackrel{}{r}`$ ลi pentru rotaลฃia finitฤ $`\alpha `$ รฎn jurul axei $`z`$.
Soluลฃie
Demonstraลฃia se obลฃine dezvoltรฎnd $`\psi (y+a))`$ รฎn serie Taylor รฎn vecinฤtatea infinitezimalฤ a punctului $`x`$, adicฤ รฎn puteri ale lui $`a`$
$$\psi (y+a)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{a^n}{n!}\frac{d^n}{dx^n}\psi (x)$$
Observฤm cฤ
$$\underset{n=0}{\overset{\mathrm{}}{}}\frac{a^n\frac{d^n}{dx^n}}{n!}=e^{a\frac{d}{dx}}$$
ลi deci $`T_a=e^{a\frac{d}{dx}}`$ รฎn cazul 1D. รn 3D, $`y=\stackrel{}{r}`$ ลi $`a\stackrel{}{a}`$. Rezultatul este $`T_\stackrel{}{a}=e^\stackrel{}{a}\stackrel{}{}`$.
Pentru rotaลฃia finitฤ $`\alpha `$ รฎn jurul lui $`z`$ avem $`y=\varphi `$ ลi $`a=\alpha `$. Rezultฤ
$$T_\alpha =R_\alpha =e^{\alpha \frac{d}{d\varphi }}.$$
O altฤ formฤ exponenลฃialฤ a rotaลฃiei รฎn $`z`$ este cea รฎn funcลฃie de operatorul moment cinetic aลa cum s-a comentat รฎn acest capitol. Fie $`x^{}=x+dx`$, ลi considerรฎnd numai primul ordin al seriei Taylor
$`\psi (x^{},y^{},z^{})`$ $`=`$ $`\psi (x,y,z)+(x^{}x){\displaystyle \frac{}{x^{}}}\psi (x^{},y^{},z^{})|_{\stackrel{}{r^{}}=\stackrel{}{r}}`$
$`+(y^{}y){\displaystyle \frac{}{y^{}}}\psi (x^{},y^{},z^{})|_{\stackrel{}{r^{}}=\stackrel{}{r}}`$
$`+(z^{}z){\displaystyle \frac{}{z^{}}}\psi (x^{},y^{},z^{})|_{\stackrel{}{r^{}}=\stackrel{}{r}}.`$
ลขinรฎnd cont de faptul cฤ
$`{\displaystyle \frac{}{x_i^{}}}\psi (\stackrel{}{r}^{})|_\stackrel{}{r}^{}`$ $`=`$ $`{\displaystyle \frac{}{x_i}}\psi (\stackrel{}{r}),`$
$`x^{}=xyd\varphi ,y^{}`$ $`=`$ $`y+xd\varphi ,z^{}=z,`$
se poate reduce seria din trei dimensiuni la numai douฤ
$`\psi (\stackrel{}{r}^{})`$ $`=`$ $`\psi (\stackrel{}{r})+(xyd\varphi x){\displaystyle \frac{\psi (\stackrel{}{r})}{x}}+(y+xd\varphi y){\displaystyle \frac{\psi (\stackrel{}{r})}{y^{}}},`$
$`=`$ $`\psi (\stackrel{}{r})yd\varphi {\displaystyle \frac{\psi (\stackrel{}{r})}{x}}+xd\varphi x{\displaystyle \frac{\psi (\stackrel{}{r})}{y}},`$
$`=`$ $`\left[1d\varphi \left(x{\displaystyle \frac{}{y}}+y{\displaystyle \frac{}{x}}\right)\right]\psi (\stackrel{}{r}).`$
Cum $`i\widehat{l}_z=\left(x\frac{}{y}y\frac{}{x}\right)`$ rezultฤ cฤ $`R=\left[1d\varphi \left(x\frac{}{y}y\frac{}{x}\right)\right].`$ Continuรฎnd รฎn ordinul doi se poate arฤta cฤ se obลฃine $`\frac{1}{2!}(i\widehat{l}_zd\varphi )^2`$ ลi aลa mai departe. Prin urmare, $`R`$ poate fi scris ca exponenลฃialฤ
$$R=e^{i\widehat{l}_zd\varphi }.$$
### Problema 3.2
Sฤ se arate cฤ pe baza expresiilor date รฎn (14) se poate ajunge la (15).
Soluลฃie
Sฤ considerฤm numai termenii lineari รฎn dezvoltarea รฎn serie Taylor (rotaลฃii infinitezimale)
$$e^{i\widehat{l}_zd\varphi }=1+i\widehat{l}_zd\varphi +\frac{1}{2!}(i\widehat{l}_zd\varphi )^2+\mathrm{},$$
ลi deci
$`(1+i\widehat{l}_zd\varphi )\widehat{A}_x(1i\widehat{l}_zd\varphi )`$ $`=`$ $`\widehat{A}_x\widehat{A}_xd\varphi ,`$
$`(\widehat{A}_x+i\widehat{l}_zd\varphi \widehat{A}_x)(1i\widehat{l}_zd\varphi )`$ $`=`$ $`\widehat{A}_x\widehat{A}_xd\varphi ,`$
$`\widehat{A}_x\widehat{A}_xi\widehat{l}_zd\varphi +i\widehat{l}_zd\varphi \widehat{A}_x+\widehat{l}_zd\varphi \widehat{A}_x\widehat{l}_zd\varphi `$ $`=`$ $`\widehat{A}_x\widehat{A}_xd\varphi ,`$
$`i(\widehat{l}_z\widehat{A}_x\widehat{A}_x\widehat{l}_z)d\varphi `$ $`=`$ $`\widehat{A}_yd\varphi .`$
Ajungem uลor la concluzia cฤ
$$[\widehat{l}_z,\widehat{A}_x]=i\widehat{A}_y.$$
Deasemenea, $`[\widehat{l}_z,\widehat{A}_y]=i\widehat{A}_x`$ se obลฃine din:
$`(1+i\widehat{l}_zd\varphi )\widehat{A}_y(1i\widehat{l}_zd\varphi )`$ $`=`$ $`\widehat{A}_xd\varphi \widehat{A}_y,`$
$`(\widehat{A}_y+i\widehat{l}_zd\varphi \widehat{A}_y)(1i\widehat{l}_zd\varphi )`$ $`=`$ $`\widehat{A}_xd\varphi \widehat{A}_y,`$
$`\widehat{A}_y\widehat{A}_yi\widehat{l}_zd\varphi +i\widehat{l}_zd\varphi \widehat{A}_y+\widehat{l}_zd\varphi \widehat{A}_y\widehat{l}_zd\varphi `$ $`=`$ $`\widehat{A}_xd\varphi \widehat{A}_y,`$
$`i(\widehat{l}_z\widehat{A}_y\widehat{A}_y\widehat{l}_z)d\varphi `$ $`=`$ $`\widehat{A}_xd\varphi .`$
### Problema 3.3
Sฤ se determine operatorul $`\frac{d\widehat{\sigma }_x}{dt}`$ pe baza Hamiltonianului unui electron cu spin aflat รฎntr-un cรฎmp magnetic de inducลฃie $`\stackrel{}{B}`$.
Soluลฃie
Hamiltonianul รฎn acest caz este $`\widehat{H}(\widehat{๐ฉ},\widehat{๐ซ},\widehat{\sigma })=\widehat{H}(\widehat{๐ฉ},\widehat{๐ซ})+\widehat{\sigma }\stackrel{}{๐}`$, unde ultimul termen este Hamiltonianul Zeeman pt. electron. Cum $`\widehat{\sigma }_x`$ comutฤ cu impulsurile ลi coordonatele, aplicarea ecuaลฃiei de miลcare Heisenberg conduce la:
$$\frac{d\widehat{\sigma }_x}{dt}=\frac{i}{\mathrm{}}[\widehat{H},\widehat{\sigma }_x]=\frac{i}{\mathrm{}}\frac{e\mathrm{}}{2m_e}((\widehat{\sigma }_yB_y+\widehat{\sigma }_zB_z)\widehat{\sigma }_x\widehat{\sigma }_x(\widehat{\sigma }_yB_y+\widehat{\sigma }_zB_z))$$
Folosind $`[\sigma _x,\sigma _y]=i\sigma _z`$, rezultฤ :
$$\frac{d\widehat{\sigma }_x}{dt}=\frac{e}{m_e}(\widehat{\sigma }_yB_z\widehat{\sigma }_zB_y)=\frac{e}{m_e}(\stackrel{}{\sigma }\times \stackrel{}{B})_x.$$
4. METODA WKB
Pentru a studia potenลฃiale mai realiste decรฎt cele de bariere ลi gropi rectangulare, este necesar de multe ori sฤ se foloseascฤ metode care sฤ permitฤ rezolvarea ecuaลฃiei Schrรถdinger pentru clase generale de potenลฃiale ลi care sฤ fie o bunฤ aproximare a soluลฃiilor exacte.
Scopul diferitelor metode de aproximare este sฤ ofere soluลฃii suficient de bune ลi simple, care sฤ permitฤ รฎn acest fel รฎnลฃelegerea comportamentului sistemului รฎn formฤ cuasianaliticฤ .
รn cadrul mecanicii cuantice, una dintre cele mai vechi ลi eficiente metode a fost dezvoltatฤ รฎn mod aproape simultan de cฤtre G. Wentzel, H. A. Kramers ลi L. Brillouin รฎn 1926, de la al cฤror nume derivฤ acronimul WKB sub care este cunoscutฤ aceastฤ metodฤ (corect este JWKB, vezi nota 4N).
Este important de menลฃionat cฤ metoda WKB se aplicฤ mai ales ecuaลฃiilor Schrรถdinger 1D ลi existฤ dificultฤลฃi serioase รฎn generalizฤrile la dimensiuni superioare.
Pentru a rezolva ecuaลฃia Schrรถdinger
$$\frac{\mathrm{}^2}{2m}\frac{d^2\psi }{dy^2}+u(y)\psi =E\psi $$
(1)
presupunem cฤ potenลฃialul are forma:
$$u(y)=u_0f\left(\frac{y}{a}\right)$$
(2)
ลi facem schimbul de variabilฤ :
$$\xi ^2=\frac{\mathrm{}^2}{2mu_0a^2}$$
(3)
$$\eta =\frac{E}{u_0}$$
(4)
$$x=\frac{y}{a}.$$
(5)
Din ecuaลฃia $`(5)`$ obลฃinem:
$$\frac{d}{dx}=\frac{dy}{dx}\frac{d}{dy}=a\frac{d}{dy}$$
(6)
$$\frac{d^2}{dx^2}=\frac{d}{dx}\left(a\frac{d}{dy}\right)=\left(a\frac{d}{dx}\right)\left(a\frac{d}{dx}\right)=a^2\frac{d^2}{dy^2}$$
(7)
ลi ec. Schrรถdinger se scrie:
$$\xi ^2\frac{d^2\psi }{dx^2}+f(x)\psi =\eta \psi .$$
(8)
Multiplicรฎnd cu $`1/\xi ^2`$ ลi definind $`r(x)=\eta f(x)`$, este posibil sฤ o scriem รฎn forma:
$$\frac{d^2\psi }{dx^2}+\frac{1}{\xi ^2}r(x)\psi =0.$$
(9)
Pentru a rezolva (9) se propune urmฤtoarea soluลฃie:
$$\psi (x)=\mathrm{exp}\left[\frac{i}{\xi }_a^xq(x)๐x\right].$$
(10)
Aลadar:
$$\frac{d^2\psi }{dx^2}=\frac{d}{dx}\left(\frac{d\psi }{x}\right)=\frac{d}{dx}\left\{\frac{i}{\xi }q(x)\mathrm{exp}\left[\frac{i}{\xi }_a^xq(x)๐x\right]\right\}$$
$$\frac{d^2\psi }{dx^2}=\frac{i}{\xi }\left\{\frac{i}{\xi }q^2(x)\mathrm{exp}\left[\frac{i}{\xi }_a^xq(x)๐x\right]+\frac{q(x)}{x}\mathrm{exp}\left[\frac{i}{\xi }_a^xq(x)๐x\right]\right\}.$$
Factorizรฎnd $`\psi `$ avem:
$$\frac{d^2\psi }{dx^2}=\left[\frac{1}{\xi ^2}q^2(x)+\frac{i}{\xi }\frac{dq(x)}{dx}\right]\psi .$$
(11)
Neglijรฎnd pe moment dependenลฃa รฎn $`x`$, ecuaลฃia Schrรถdinger se poate scrie :
$$\left[\frac{1}{\xi ^2}q^2+\frac{i}{\xi }\frac{q}{x}+\frac{1}{\xi ^2}r\right]\psi =0$$
(12)
ลi cum รฎn general $`\psi 0`$, avem:
$$i\xi \frac{dq}{dx}+rq^2=0,$$
(13)
care este o ecuaลฃie diferenลฃialฤ linearฤ de tip Riccati, a cฤrei soluลฃii se cautฤ รฎn forma unei serii de puteri ale lui $`\xi `$ cu presupunerea cฤ $`\xi `$ este foarte mic.
Mai exact seria se propune de forma:
$$q(x)=\underset{n=0}{\overset{\mathrm{}}{}}(i\xi )^nq_n(x).$$
(14)
Substituind-o รฎn Riccati obลฃinem:
$$i\xi \underset{n=0}{\overset{\mathrm{}}{}}(i\xi )^n\frac{dq_n}{dx}+r(x)\underset{\mu =0}{\overset{\mathrm{}}{}}(i\xi )^\mu q_\mu \underset{\nu =0}{\overset{\mathrm{}}{}}(i\xi )^\nu q_\nu =0.$$
(15)
Printr-o rearanjare a termenilor obลฃinem:
$$\underset{n=0}{\overset{\mathrm{}}{}}(1)^n(i\xi )^{n+1}\frac{dq_n}{dx}+r(x)\underset{\mu =0}{\overset{\mathrm{}}{}}\underset{\nu =0}{\overset{\mathrm{}}{}}(i\xi )^{\mu +\nu }q_\mu q_\nu =0.$$
(16)
Seriile duble au urmฤtoarea proprietate:
$$\underset{\mu =0}{\overset{\mathrm{}}{}}\underset{\nu =0}{\overset{\mathrm{}}{}}a_{\mu \nu }=\underset{n=0}{\overset{\mathrm{}}{}}\underset{m=0}{\overset{n}{}}a_{m,nm},$$
unde: $`\mu =nm,\nu =m`$ .
รn acest fel:
$$\underset{n=0}{\overset{\mathrm{}}{}}(1)^n(i\xi )^{n+1}\frac{dq_n}{dx}+r(x)\underset{n=0}{\overset{\mathrm{}}{}}\underset{m=0}{\overset{n}{}}(i\xi )^{nm+m}q_mq_{nm}=0.$$
(17)
Sฤ vedem explicit cรฎลฃiva termeni รฎn fiecare din seriile din ecuaลฃia (17):
$$\underset{n=0}{\overset{\mathrm{}}{}}(1)^n(i\xi )^{n+1}\frac{dq_n}{dx}=i\xi \frac{dq_0}{dx}+\xi ^2\frac{dq_1}{dx}i\xi ^3\frac{dq_2}{dx}+\mathrm{}$$
(18)
$$\underset{n=0}{\overset{\mathrm{}}{}}\underset{m=0}{\overset{n}{}}(i\xi )^nq_mq_{nm}=q_0^2i2\xi q_0q_1+\mathrm{}$$
(19)
Pentru ca ambele serii sฤ conลฃinฤ pe $`i\xi `$ รฎn primul termen trebuie sฤ le scriem:
$$\underset{n=1}{\overset{\mathrm{}}{}}(1)^{n1}(i\xi )^n\frac{dq_{n1}}{dx}+r(x)q_0^2\underset{n=1}{\overset{\mathrm{}}{}}\underset{m=0}{\overset{n}{}}(i\xi )^nq_mqnm=0,$$
ceea ce ne conduce la:
$$\underset{n=1}{\overset{\mathrm{}}{}}\left[(i\xi )^n\frac{dq_{n1}}{dx}\underset{m=0}{\overset{n}{}}(i\xi )^nq_mq_{nm}\right]+\left[r(x)q_0^2\right]=0.$$
(20)
Pentru ca aceastฤ ecuaลฃie sฤ se satisfacฤ avem condiลฃiile:
$$r(x)q_0^2=0q_0=\pm \sqrt{r(x)}$$
(21)
$$(i\xi )^n\frac{dq_{n1}}{dx}\underset{m=0}{\overset{n}{}}(i\xi )^nq_mq_{nm}=0$$
$$\frac{dq_{n1}}{dx}=\underset{m=0}{\overset{n}{}}q_mq_{nm}n1.$$
(22)
Aceasta ultima este o relaลฃie de recurenลฃฤ care apare รฎn metoda WKB. Este momentul sฤ amintim cฤ am definit $`r(x)=\eta f(x),\eta =\frac{E}{u_0}\&f(x)=\frac{u}{u_0}`$ ลi cu ajutorul ec. $`(21)`$ obลฃinem:
$$q_0=\pm \sqrt{\eta f(x)}=\pm \sqrt{\frac{E}{u_0}\frac{u}{u_0}}=\pm \sqrt{\frac{2m(Eu)}{2mu_0}},$$
(23)
care ne indicฤ natura clasicฤ a impulsului WKB a particulei de energie $`E`$ รฎn potenลฃialul $`u`$ ลi รฎn unitฤลฃi $`\sqrt{2mu_0}`$. Astfel:
$$q_0=p(x)=\sqrt{\eta f(x)}$$
nu este un operator. Dacฤ aproximฤm pรฎnฤ la ordinul doi, obลฃinem:
$$q(x)=q_0i\xi q_1\xi ^2q_2$$
ลi folosind relaลฃia de recurenลฃฤ WKB (22) calculฤm $`q_1`$ ลi $`q_2`$:
$$\frac{dq_0}{dx}=2q_0q_1q_1=\frac{1}{2}\frac{\frac{dq_0}{dx}}{q_0}=\frac{1}{2}\frac{d}{dx}(\mathrm{ln}|q_0|)$$
$$q_1=\frac{1}{2}\frac{d}{dx}(\mathrm{ln}|p(x)|)$$
(24)
$$\frac{dq_1}{dx}=2q_0q_2q_1^2q_2=\frac{\frac{dq_1}{dx}q_1^2}{2q_0}.$$
(25)
Din ecuaลฃia $`(24)`$, ne dฤm seama cฤ mฤrimea $`q_1`$ este panta cu semnul schimbat a lui $`\mathrm{ln}|q_0|`$; cรฎnd $`q_0`$ este foarte mic, $`q_10\xi q_10`$ ลi prin urmare seria diverge. Pentru a evita acest lucru se impune urmฤtoarea condiลฃie WKB:
$$|q_0||\xi q_1|=\xi |q_1|.$$
Este important de observat cฤ aceastฤ condiลฃie WKB nu se satisface pentru acele puncte $`x_k`$ unde:
$$q_0(x_k)=p(x_k)=0.$$
Dar $`q_0=p=\sqrt{\frac{2m(Eu)}{2mu_0}}`$ ลi deci ecuaลฃia precedentฤ ne conduce la:
$$E=u(x_k).$$
(26)
รn mecanica clasicฤ punctele $`x_k`$ care satisfac (26) se numesc puncte de รฎntoarcere pentru cฤ รฎn ele are loc schimbarea sensului de miลcare a particulei macroscopice.
รn baza acestor argumente, putem sฤ spunem despre $`q_0`$ cฤ este o soluลฃie clasicฤ a problemei examinate ลi cฤ mฤrimile $`q_1`$ & $`q_2`$ sunt respectiv prima ลi a doua corecลฃie cuanticฤ รฎn problema WKB.
Pentru a obลฃine funcลฃiile de undฤ vom considera numai soluลฃia clasicฤ ลi prima corecลฃie cuanticฤ a problemei pe care le substituim รฎn forma WKB a lui $`\psi `$:
$$\psi =\mathrm{exp}\left[\frac{i}{\xi }_a^xq(x)๐x\right]=\mathrm{exp}\left[\frac{i}{\xi }_a^x(q_0i\xi q_1)๐x\right]$$
$$\psi =\mathrm{exp}\left(\frac{i}{\xi }_a^xq_0๐x\right)\mathrm{exp}\left(_a^xq_1๐x\right).$$
Pentru al doilea factor avem:
$$\mathrm{exp}\left(_a^xq_1๐x\right)=\mathrm{exp}\left[\frac{1}{2}_a^x\frac{d}{dx}(\mathrm{ln}|p(x)|)๐x\right]=$$
$$=\mathrm{exp}\left[\frac{1}{2}(\mathrm{ln}|p(x)|)|_a^x\right]=\frac{A}{\sqrt{p(x)}},$$
cu $`A`$ o constantฤ , รฎn timp ce pentru primul factor se obลฃine:
$$\mathrm{exp}\left(\frac{i}{\xi }_a^xq_0๐x\right)=\mathrm{exp}\left[\pm \frac{i}{\xi }_a^xp(x)๐x\right].$$
Astfel putem scrie $`\psi `$ รฎn urmฤtoarea formฤ :
$$\psi ^\pm =\frac{1}{\sqrt{p(x)}}\mathrm{exp}\left[\pm \frac{i}{\xi }_a^xp(x)๐x\right],$$
(27)
care sunt cunoscute ca soluลฃii WKB ale ecuaลฃiei Schrรถdinger unidimensionale. Soluลฃia generalฤ WKB รฎn regiunea รฎn care condiลฃia WKB se satisface se scrie:
$$\psi =a_+\psi ^++a_{}\psi ^{}.$$
(28)
Aลa cum deja s-a menลฃionat, nu existฤ soluลฃie WKB รฎn punctele de รฎntoarcere, ceea ce ridicฤ problema modului รฎn care se face trecerea de la $`\psi (x<x_k)`$ la $`\psi (x>x_k)`$. Rezolvarea acestei dificultฤลฃi se face prin introducerea formulelor de conexiune WKB.
### Formulele de Conexiune
Deja s-a vฤzut cฤ soluลฃiile WKB sunt singulare รฎn punctele de รฎntoarcere clasicฤ ; totuลi aceste soluลฃii sunt corecte la stรฎnga ลi la dreapta acestor puncte $`x_k`$. Ne รฎntrebฤm deci cum schimbฤm $`\psi (x<x_k)`$ รฎn $`\psi (x>x_k)`$ รฎn aceste puncte. Rฤspunsul este dat de aลa numitele formule de conexiune.
Din teoria ecuaลฃiilor diferenลฃiale ordinare ลi pe baza analizei de funcลฃii de variabilฤ complexฤ se poate demonstra cฤ formulele de conexiune existฤ ลi cฤ sunt urmฤtoarele:
$$\psi _1(x)=\frac{1}{\left[r(x)\right]^{\frac{1}{4}}}\mathrm{exp}\left(_x^{x_k}\sqrt{r(x)}๐x\right)$$
$$\frac{2}{\left[r(x)\right]^{\frac{1}{4}}}\mathrm{cos}\left(_{x_k}^x\sqrt{r(x)}๐x\frac{\pi }{4}\right),$$
(29)
unde $`\psi _1(x)`$ are numai comportament atenuant exponenลฃial pentru $`x<x_k`$. Prima formulฤ de conexiune aratฤ cฤ funcลฃia $`\psi (x)`$, care la stรฎnga punctului de รฎntoarcere se comportฤ atenuant exponenลฃial, trece รฎn dreapta acelui punct รฎntr-o cosinusoidฤ de fazฤ $`\varphi =\frac{\pi }{4}`$ ลi de amplitudine dublฤ faลฃฤ de amplitudinea exponenลฃialei.
รn cazul unei funcลฃii $`\psi (x)`$ mai generale, respectiv o funcลฃie care sฤ aibฤ un comportament exponenลฃial crescฤtor ลi atenuant, formula de conexiune corespunzฤtoare este:
$$\mathrm{sin}\left(\varphi +\frac{\pi }{4}\right)\frac{1}{\left[r(x)\right]^{\frac{1}{4}}}\mathrm{exp}\left(_x^{x_k}\sqrt{r(x)}๐x\right)$$
$$\frac{1}{\left[r(x)\right]^{\frac{1}{4}}}\mathrm{cos}\left(_{x_k}^x\sqrt{r(x)}๐x+\varphi \right),$$
(30)
cu condiลฃia ca $`\varphi `$ sฤ nu ia o valoare prea apropiatฤ de $`\frac{\pi }{4}`$. Motivul este cฤ dacฤ $`\varphi =\frac{\pi }{4}`$, funcลฃia sinus se anuleazฤ . Aceastฤ a doua formulฤ de conexiune significฤ cฤ o funcลฃie care se comportฤ ca o cosinusoidฤ la dreapta unui punct de รฎntoarcere trece รฎn partea sa stรฎnga ca o exponenลฃialฤ crescฤtoare cu amplitudinea modulatฤ de cฤtre o sinusoidฤ .
Pentru a studia detaliile procedurii de obลฃinere a formulelor de conexiune se poate consulta expunerea din cartea Mathematical Methods of Physics de J. Mathews & R.L. Walker.
### Estimarea erorii introduse de aproximaลฃia WKB
Am gฤsit soluลฃia ecuaลฃiei Schrรถdinger รฎn orice regiune unde se satisface condiลฃia WKB. Totuลi, soluลฃiile WKB diverg รฎn punctele de รฎntoarcere aลa cum am semnalat. Vom analiza, deลi superficial, aceastฤ problematicฤ cu scopul de a propune formulele de conexiune รฎntr-o vecinฤtate redusฤ a punctelor de รฎntoarcere.
Sฤ presupunem cฤ $`x=x_k`$ este un punct de รฎntoarcere, unde avem: $`q_0(x_k)=p(x_k)=0E=u(x_k)`$. La stรฎnga lui $`x_k`$, adicฤ รฎn semidreapta $`x<x_k`$, vom presupune cฤ $`E<u(x)`$, astfel cฤ รฎn aceastฤ regiune soluลฃia WKB este:
$$\psi (x)=\frac{a}{\left[\frac{u(x)E}{u_0}\right]^{\frac{1}{4}}}\mathrm{exp}\left(\frac{1}{\xi }_x^{x_k}\sqrt{\frac{u(x)E}{u_0}}๐x\right)+$$
$$+\frac{b}{\left[\frac{u(x)E}{u_0}\right]^{\frac{1}{4}}}\mathrm{exp}\left(\frac{1}{\xi }_x^{x_k}\sqrt{\frac{u(x)E}{u_0}}๐x\right).$$
(31)
รn acelaลi mod, la dreapta lui $`x_k`$ ( รฎn semidreapta $`x>x_k`$ ) presupunem $`E>u(x)`$. รn consecinลฃฤ soluลฃia WKB รฎn aceastฤ zonฤ este:
$$\psi (x)=\frac{c}{\left[\frac{Eu(x)}{u_0}\right]^{\frac{1}{4}}}\mathrm{exp}\left(\frac{i}{\xi }_{x_k}^x\sqrt{\frac{Eu(x)}{u_0}}๐x\right)+$$
$$+\frac{d}{\left[\frac{Eu(x)}{u_0}\right]^{\frac{1}{4}}}\mathrm{exp}\left(\frac{i}{\xi }_{x_k}^x\sqrt{\frac{Eu(x)}{u_0}}๐x\right).$$
(32)
Dacฤ $`\psi (x)`$ este o funcลฃie realฤ , va avea aceastฤ proprietate atรฎt la dreapta cรฎt ลi la stรฎnga lui $`x_k`$. Vom numi acest fapt โcondiลฃia de realitateโ, care รฎnseamnฤ cฤ dacฤ $`a,b\mathrm{}`$, atunci $`c=d^{}`$.
Problema noastrฤ este de a conecta aproximaลฃiile din cele douฤ laturi ale lui $`x_k`$ pentru ca ele sฤ se refere la aceeaลi soluลฃie. Aceasta รฎnseamnฤ a gฤsi $`c`$ ลi $`d`$ dacฤ se cunosc $`a`$ ลi $`b`$, precum ลi viceversa. Pentru a efectua aceastฤ conexiune, trebuie sฤ utilizฤm o soluลฃie aproximatฤ , care sฤ fie corectฤ de-a lungul unui drum care leagฤ regiunile din cele douฤ laturi ale lui $`x_k`$, unde soluลฃiile WKB sฤ fie deasemenea corecte.
Cel mai comun este sฤ se recurgฤ la o metodฤ propusฤ de cฤtre Zwann ลi Kemble care constฤ รฎn a ieลi de pe axa realฤ รฎn vecinฤtatea lui $`x_k`$, pe un contur รฎn jurul lui $`x_k`$ รฎn planul complex. Pe acest contur se considerฤ cฤ soluลฃiile WKB continuฤ sฤ fie corecte. รn aceastฤ prezentare vom folosi aceastฤ metodฤ , dar numai cu scopul de a obลฃine un mijloc de a estima erorile รฎn aproximaลฃia WKB.
Estimarea erorilor este รฎntotdeauna importantฤ pentru soluลฃiile aproximate prin diferite metode ลi รฎn plus รฎn cazul WKB aproximaลฃia se face pe intervale mari ale axei reale, ceea ce poate duce la acumularea erorilor ลi la eventuale artefacte datorate ลifturilor de fazฤ astfel introduse.
Sฤ definim funcลฃiile WKB asociate รฎn felul urmฤtor:
$$W_\pm =\frac{1}{\left[\frac{Eu(x)}{u_0}\right]^{\frac{1}{4}}}\mathrm{exp}\left(\pm \frac{i}{\xi }_{x_k}^x\sqrt{\frac{Eu(x)}{u_0}}๐x\right),$$
(33)
pe care le considerฤm ca funcลฃii de variabilฤ complexฤ . Vom folosi tฤieturi pentru a evita discontinuitฤลฃile din zerourile lui $`r(x)=\frac{Eu(x)}{u_0}`$. Aceste funcลฃii satisfac o ecuaลฃie diferenลฃialฤ care se poate obลฃine diferenลฃiindu-le รฎn raport cu $`x`$, conducรฎnd la:
$$W_\pm ^{}=\left(\pm \frac{i}{\xi }\sqrt{r}\frac{1}{4}\frac{r^{}}{r}\right)W_\pm $$
$$W_\pm ^{\prime \prime }+\left[\frac{r}{\xi ^2}+\frac{1}{4}\frac{r^{\prime \prime }}{r}\frac{5}{16}\left(\frac{r^{}}{r}\right)^2\right]W_\pm =0.$$
(34)
Sฤ notฤm:
$$s(x)=\frac{1}{4}\frac{r^{\prime \prime }}{r}\frac{5}{16}\left(\frac{r^{}}{r}\right)^2,$$
(35)
atunci $`W_\pm `$ sunt soluลฃii exacte ale ecuaลฃiei:
$$W_\pm ^{\prime \prime }+\left[\frac{1}{\xi ^2}r(x)+s(x)\right]W_\pm =0,$$
(36)
dar satisfac numai aproximativ ecuaลฃia Schrรถdinger, care este regularฤ รฎn $`x=x_k`$ รฎn timp ce ecuaลฃia pentru funcลฃiile WKB asociate este singularฤ รฎn acel punct.
Vom defini funcลฃiile $`\alpha _\pm (x)`$ care sฤ satisfacฤ urmฤtoarele douฤ relaลฃii:
$$\psi (x)=\alpha _+(x)W_+(x)+\alpha _{}(x)W_{}(x)$$
(37)
$$\psi ^{}(x)=\alpha _+(x)W_+^{}(x)+\alpha _{}(x)W_{}^{}(x),$$
(38)
unde $`\psi (x)`$ este soluลฃie a ecuaลฃiei Schrรถdinger. Rezolvรฎnd ecuaลฃiile anterioare pentru $`\alpha _\pm `$ avem:
$$\alpha _+=\frac{\psi W_{}^{}\psi ^{}W_{}}{W_+W_{}^{}W_+^{}W_{}}\alpha _{}=\frac{\psi W_+^{}\psi ^{}W_+}{W_+W_{}^{}W_+^{}W_{}},$$
unde numฤrฤtorul este exact Wronskianul lui $`W_+`$ ลi $`W_{}`$. Nu este dificil de demonstrat cฤ acesta ia valoarea $`\frac{2}{\xi }i`$, astfel cฤ $`\alpha _\pm `$ se simplificฤ la forma:
$$\alpha _+=\frac{\xi }{2}i\left(\psi W_{}^{}\psi ^{}W_{}\right)$$
(39)
$$\alpha _{}=\frac{\xi }{2}i\left(\psi W_+^{}\psi ^{}W_+\right).$$
(40)
Efectuรฎnd derivata รฎn $`x`$ รฎn ecuaลฃiile $`(39)`$ ลi $`(40)`$, avem:
$$\frac{d\alpha _\pm }{dx}=\frac{\xi }{2}i\left(\psi ^{}W_{}^{}+\psi W_{}^{\prime \prime }\psi ^{\prime \prime }W_{}\psi ^{}W_{}^{}\right).$$
(41)
รn paranteze, primul ลi al patrulea termen se anuleazฤ ; amintim cฤ :
$$\psi ^{\prime \prime }+\frac{1}{\xi ^2}r(x)\psi =0\&W_\pm ^{\prime \prime }+\left[\frac{1}{\xi ^2}r(x)+s(x)\right]W_\pm =0$$
ลi deci putem scrie ecuaลฃia $`(41)`$ รฎn forma:
$$\frac{d\alpha _\pm }{dx}=\frac{\xi }{2}i\left[\psi \left(\frac{r}{\xi ^2}+s\right)W_{}+\frac{r}{\xi ^2}\psi W_{}\right]$$
$$\frac{d\alpha _\pm }{dx}=\frac{\xi }{2}is(x)\psi (x)W_{}(x),$$
(42)
care รฎn baza ecuaลฃiilor $`(33)`$ ลi $`(37)`$ devine:
$$\frac{d\alpha _\pm }{dx}=\frac{\xi }{2}i\frac{s(x)}{\left[r(x)\right]^{\frac{1}{2}}}\left[\alpha _\pm +\alpha _{}\mathrm{exp}\left(\frac{2}{\xi }i_{x_k}^x\sqrt{r(x)}๐x\right)\right].$$
(43)
Ecuaลฃiile $`(42)`$ ลi $`(43)`$ se folosesc pentru a estima eroarea care se comite รฎn aproximaลฃia WKB รฎn cazul unidimensional.
Motivul pentru care $`\frac{d\alpha _\pm }{dx}`$ se poate considera ca mฤsurฤ a erorii WKB este cฤ รฎn ecuaลฃiile $`(31)`$ ลi $`(32)`$ constantele $`a`$, $`b`$ ลi $`c`$, $`d`$, respectiv, ne dau numai soluลฃii $`\psi `$ aproximative, รฎn timp ce funcลฃiile $`\alpha _\pm `$ introduse รฎn ecuaลฃiile $`(37)`$ ลi $`(38)`$ produc soluลฃii $`\psi `$ exacte. Din punct de vedere geometric derivata dฤ panta dreptei tangente la aceste funcลฃii ลi indicฤ mฤsura รฎn care $`\alpha _\pm `$ deviazฤ faลฃฤ de constantele $`a`$, $`b`$, $`c`$ ลi $`d`$.
4N. Notฤ: Articolele (J)WKB originale sunt urmฤtoarele:
G. Wentzel, โEine Verallgemeinerung der Wellenmechanikโ, \[โO generalizare a mecanicii ondulatoriiโ\],
Zeitschrift fรผr Physik 38, 518-529 (1926) \[primit 18 June 1926\]
L. Brillouin, โLa mรฉcanique ondulatoire de Schrรถdinger: une mรฉthode gรฉnรฉrale de resolution par approximations successivesโ, \[โMecanica ondulatorie a lui Schrรถdinger: o metodฤ generalฤ de rezolvare prin aproximฤri succesiveโ\],
Comptes Rendus Acad. Sci. Paris 183, 24-26 (1926) \[primit 5 July 1926\]
H.A. Kramers, โWellenmechanik und halbzahlige Quantisierungโ, \[โMecanica ondulatorie ลi cuantizarea semiรฎntreagฤ โ\],
Zf. Physik 39, 828-840 (1926) \[primit 9 Sept. 1926\]
H. Jeffreys, โOn certain approx. solutions of linear diff. eqs. of the second orderโ, \[โAsupra unor soluลฃii aproximative a ecuaลฃiilor diferenลฃiale lineare de ordinul doiโ\],
Proc. Lond. Math. Soc. 23, 428-436 (1925)
4P. Probleme
Problema 4.1
Sฤ se foloseascฤ metoda WKB pentru o particulฤ de energie $`E`$ care se miลcฤ รฎntr-un potenลฃial $`u(x)`$ de forma arฤtatฤ รฎn figura 4.1.
Soluลฃie
Ecuaลฃia Schrรถdinger corespunzฤtoare este:
$$\frac{d^2\psi }{dx^2}+\frac{2m}{\mathrm{}^2}\left[Eu(x)\right]\psi =0.$$
(44)
Dupฤ cum putem vedea:
$$r(x)=\frac{2m}{\mathrm{}^2}\left[Eu(x)\right]\{\begin{array}{cc}\text{este pozitivฤ pentru }a<x<b\hfill & \\ \text{este negativฤ pentru }x<a,x>b\text{.}\hfill & \end{array}$$
Dacฤ $`\psi (x)`$ corespunde zonei รฎn care $`x<a`$, la trecerea รฎn intervalul $`a<x<b`$, formula de conexiune este datฤ de ecuaลฃia $`(29)`$ ลi ne spune cฤ :
$$\psi (x)\frac{A}{\left[Eu\right]^{\frac{1}{4}}}\mathrm{cos}\left(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)$$
(45)
unde $`A`$ este o constantฤ arbitrarฤ .
Cรฎnd $`\psi (x)`$ corespunde zonei $`x>b`$, la trecerea รฎn intervalul $`a<x<b`$ avem รฎn mod similar:
$$\psi (x)\frac{B}{\left[Eu\right]^{\frac{1}{4}}}\mathrm{cos}\left(_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right),$$
(46)
unde $`B`$ este o constantฤ arbitrarฤ . Motivul pentru care formula de conexiune este din nou ecuaลฃia (29), se รฎnลฃelege examinรฎnd ce se รฎntรฎmplฤ cรฎnd particula ajunge la al doilea punct clasic de รฎntoarcere $`x=b`$. Acesta produce inversia direcลฃiei de miลcare ลi atunci particula apare ca venind de la dreapta spre stรฎnga. Cu alte cuvinte, ne gฤsim รฎn prima situaลฃie (de la stรฎnga la dreapta) numai cฤ vฤzutฤ รฎntr-o oglindฤ รฎn punctul $`x=a`$.
Aceste douฤ expresii trebuie sฤ fie aceleaลi independent de constantele $`A`$ ลi $`B`$, astfel cฤ :
$$\mathrm{cos}\left(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)=\mathrm{cos}\left(_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)$$
$$\mathrm{cos}\left(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)+\mathrm{cos}\left(_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)=0.$$
(47)
Amintind cฤ :
$$\mathrm{cos}A+\mathrm{cos}B=2\mathrm{cos}\left(\frac{A+B}{2}\right)\mathrm{cos}\left(\frac{AB}{2}\right),$$
ecuaลฃia $`(47)`$ se scrie:
$$2\mathrm{cos}\left[\frac{1}{2}(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}dx\frac{\pi }{4}+_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}dx\frac{\pi }{4})\right]$$
$$\mathrm{cos}\left[\frac{1}{2}(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}dx\frac{\pi }{4}_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}dx+\frac{\pi }{4})\right]=0,$$
(48)
ceea ce implicฤ pentru argumentele acestor cosinusoide cฤ sunt multipli รฎntregi de $`\frac{\pi }{2}`$; argumentul celui de-a doua cosinusoide nu ne duce la nici un rezultat netrivial, aลa cฤ ne fixฤm atenลฃia numai asupra argumentului primei cosinusoide, care este esenลฃial pentru obลฃinerea unui rezultat important:
$$\frac{1}{2}\left(_a^x\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}+_x^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{4}\right)=\frac{n}{2}\pi \text{pt. n impar}$$
$$_a^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x\frac{\pi }{2}=n\pi $$
$$_a^b\sqrt{\frac{2m}{\mathrm{}^2}(Eu)}๐x=(n+\frac{1}{2})\pi $$
$$_a^b\sqrt{2m(Eu)}๐x=(n+\frac{1}{2})\pi \mathrm{}.$$
(49)
Acest rezultat este foarte similar regulilor de cuantizare Bohr - Sommerfeld.
Amintim cฤ postulatul lui Bohr stabileลte cฤ momentul unghiular al unui electron care se miลcฤ pe o โorbitฤ permisฤ โ รฎn jurul nucleului atomic este cuantizatฤ ลi valoarea sa este: $`L=n\mathrm{}`$, $`n=1,2,3,\mathrm{}`$. Amintim deasemenea cฤ regulile de cuantizare Wilson - Sommerfeld stabilesc cฤ orice coordonatฤ a unui sistem fizic care variazฤ periodic รฎn timp trebuie sฤ satisfacฤ condiลฃia cuanticฤ : $`p_q๐q=n_qh`$; unde $`q`$ este o coordonatฤ periodicฤ , $`p_q`$ este impulsul asociat acesteia, $`n_q`$ este un numฤr รฎntreg ลi $`h`$ este constanta lui Planck. Se vede cฤ rezultatul obลฃinut รฎn aproximaลฃia WKB este cu adevฤrat foarte asemฤnฤtor.
Problema 4.2
Sฤ se estimeze eroarea care se comite รฎn soluลฃia WKB รฎntr-un punct $`x_1x_k`$, cu $`x_k`$ un punct clasic de รฎntoarcere, pentru ecuaลฃia diferenลฃialฤ $`y^{\prime \prime }+xy=0`$. Soluลฃia acestei probleme este importantฤ รฎn studiul cรฎmpurilor uniforme, aลa cum sunt cele gravitaลฃionale sau cele electrice produse de plฤci plane รฎncฤrcate.
Soluลฃie:
Pentru aceastฤ ecuaลฃie diferenลฃialฤ avem:
$$\xi =1,r(x)=x\&s(x)=\frac{5}{16}x^2.$$
$`r(x)=x`$ are un singur zero รฎn $`x_k=0`$, astfel cฤ pentru $`x0`$:
$$W_\pm =x^{\frac{1}{4}}\mathrm{exp}\left(\pm i_0^x\sqrt{x}๐x\right)=x^{\frac{1}{4}}\mathrm{exp}\left(\pm \frac{2}{3}ix^{\frac{3}{2}}\right).$$
(50)
Derivรฎnd $`W_\pm `$ pรฎnฤ la a doua derivatฤ รฎn $`x`$, ne dฤm seama cฤ se satisface urmฤtoarea ecuaลฃie diferenลฃialฤ :
$$W_\pm ^{\prime \prime }+(x\frac{5}{16}x^2)W_\pm =0.$$
(51)
Soluลฃia exactฤ $`y(x)`$ a acestei ecuaลฃii diferenลฃiale o scriem ca o combinaลฃie linearฤ de $`W_\pm `$, aลa cum s-a indicat รฎn secลฃiunea corespondentฤ estimฤrii erorii รฎn aproximaลฃia WKB; amintim cฤ s-a propus o combinaลฃie linearฤ de forma:
$$y(x)=\alpha _+(x)W_+(x)+\alpha _{}(x)W_{}(x)$$
Pentru $`x`$ mari, o soluลฃie generalฤ a ecuaลฃiei noastre diferenลฃiale se poate scrie รฎn aproximaลฃia WKB รฎn forma:
$$y(x)=Ax^{\frac{1}{4}}\mathrm{cos}\left(\frac{2}{3}x^{\frac{3}{2}}+\delta \right)\text{cรฎnd}x\mathrm{},$$
(52)
astfel cฤ $`\alpha _+\frac{A}{2}e^{i\delta }`$ ลi $`\alpha _{}\frac{A}{2}e^{i\delta }`$ pentru $`x\mathrm{}`$. Vrem sฤ calculฤm eroarea datoratฤ acestei soluลฃii WKB. O mฤsurฤ a acestei erori este devierea lui $`\alpha _+`$ ลi a lui $`\alpha _{}`$ faลฃฤ de constantele $`A`$. Pentru aceasta folosim ecuaลฃia:
$$\frac{d\alpha _\pm }{dx}=\frac{\xi }{2}i\frac{s(x)}{\sqrt{r(x)}}\left[\alpha _\pm +\alpha _{}\mathrm{exp}\left(2i_{x_k}^x\sqrt{r(x)}๐x\right)\right]$$
ลi efectuรฎnd substituลฃiile corespunzฤtoare avem:
$$\frac{d\alpha _\pm }{dx}=\frac{i}{2}\left(\frac{5}{16}x^2\right)x^{\frac{1}{2}}\left[\frac{A}{2}e^{\pm i\delta }+\frac{A}{2}e^{i\delta }\mathrm{exp}\left(2i\frac{2}{3}x^{\frac{3}{2}}\right)\right].$$
(53)
ลtim cฤ $`\mathrm{\Delta }\alpha _\pm `$ reprezintฤ schimbฤrile pe care le prezintฤ $`\alpha _\pm `$ cรฎnd $`x`$ variazฤ รฎntre $`x_1`$ ลi $`\mathrm{}`$, ceea ce ne permite calculul prin intermediul lui:
$$\frac{\mathrm{\Delta }\alpha _\pm }{A/2}=\frac{2}{A}_{x_1}^{\mathrm{}}\frac{d\alpha _\pm }{dx}๐x=$$
$$=\pm i\frac{5}{32}e^{\pm i\delta }\left[\frac{2}{3}x_1^{\frac{3}{2}}+e^{2i\delta }_{x_1}^{\mathrm{}}x^{\frac{5}{2}}\mathrm{exp}\left(i\frac{4}{3}x^{\frac{3}{2}}\right)๐x\right].$$
(54)
Al doilea termen din paranteze este mai puลฃin important decรฎt primul pentru cฤ exponenลฃiala complexฤ oscileazฤ รฎntre $`1`$ ลi $`1`$ ลi deci $`x^{\frac{5}{2}}<x^{\frac{3}{2}}`$. Prin urmare:
$$\frac{\mathrm{\Delta }\alpha _\pm }{A/2}\pm \frac{5}{48}ie^{\pm i\delta }x_1^{\frac{3}{2}}$$
(55)
ลi cum putem vedea eroarea care se introduce este รฎntr-adevฤr micฤ dacฤ tinem cont ลi de faptul cฤ exponenลฃiala complexฤ oscileazฤ รฎntre $`1`$ ลi $`1`$, iar $`x_1^{\frac{3}{2}}`$ este deasemenea mic.
## 5. OSCILATORUL ARMONIC (OA)
## Soluลฃia ecuaลฃiei Schrรถdinger pentru OA
Oscilatorul armonic (OA) poate fi considerat ca o paradigmฤ a Fizicii. Utilitatea sa apare รฎn marea majoritate a domeniilor, de la fizica clasicฤ pรฎnฤ la electrodinamica cuanticฤ ลi teorii ale obiectelor colapsate gravitaลฃional.
Din mecanica clasicฤ ลtim cฤ multe potenลฃiale complicate pot fi aproximate รฎn vecinฤtatea poziลฃiilor de echilibru printr-un potenลฃial OA
$$V(x)\frac{1}{2}V^{\prime \prime }(a)(xa)^2.$$
(1)
Acesta este un caz unidimensional. Pentru acest caz, funcลฃia Hamiltonianฤ clasicฤ a unei particule de masฤ m, oscilรฎnd cu frecvenลฃa $`\omega `$ are urmฤtoarea formฤ :
$$H=\frac{p^2}{2m}+\frac{1}{2}m\omega ^2x^2$$
(2)
ลi Hamiltonianul cuantic corespunzฤtor รฎn spaลฃiul de configuraลฃii este :
$$\widehat{H}=\frac{1}{2m}(i\mathrm{}\frac{d}{dx})^2+\frac{1}{2}m\omega ^2x^2$$
(3)
$$\widehat{H}=\frac{\mathrm{}^2}{2m}\frac{d^2}{dx^2}+\frac{1}{2}m\omega ^2x^2.$$
(4)
Dat faptul cฤ potenลฃialul este independent de timp, FP $`\mathrm{\Psi }_n`$ ลi autovalorile $`E_n`$ se determinฤ cu ajutorul ecuaลฃiei Schrรถdinger independentฤ de timp :
$$\widehat{H}\mathrm{\Psi }_n=E_n\mathrm{\Psi }_n.$$
(5)
Considerรฎnd Hamiltonianul pentru OA , ecuaลฃia Schrรถdinger pentru acest caz este :
$$\frac{d^2\mathrm{\Psi }}{dx^2}+\left[\frac{2mE}{\mathrm{}^2}\frac{m^2\omega ^2}{\mathrm{}^2}x^2\right]\mathrm{\Psi }=0.$$
(6)
Am suprimat subindicii lui $`E`$ ลi $`\mathrm{\Psi }`$ pentru cฤ nu au nici o importanลฃฤ aici. Definind:
$$k^2=\frac{2mE}{\mathrm{}^2}$$
(7)
$$\lambda =\frac{m\omega }{\mathrm{}},$$
(8)
ecuaลฃia Schrรถdinger devine:
$$\frac{d^2\mathrm{\Psi }}{dx^2}+[k^2\lambda ^2x^2]\mathrm{\Psi }=0,$$
(9)
cunoscutฤ ca ecuaลฃia diferenลฃialฤ Weber รฎn matematicฤ .
Vom face รฎn continuare transformarea:
$$y=\lambda x^2.$$
(10)
รn general, cu schimbul de variabilฤ de la $`x`$ la $`y`$ , operatorii diferenลฃiali iau forma:
$$\frac{d}{dx}=\frac{dy}{dx}\frac{d}{dy}$$
(11)
$$\frac{d^2}{dx^2}=\frac{d}{dx}(\frac{dy}{dx}\frac{d}{dy})=\frac{d^2y}{dx^2}\frac{d}{dy}+(\frac{dy}{dx})^2\frac{d^2}{dy^2}.$$
(12)
Aplicรฎnd aceastฤ regulฤ evidentฤ transformฤrii propuse obลฃinem urmฤtoarea ecuaลฃie diferenลฃialฤ รฎn variabila $`y`$ :
$$y\frac{d^2\mathrm{\Psi }}{dy^2}+\frac{1}{2}\frac{d\mathrm{\Psi }}{dy}+[\frac{k^2}{4\lambda }\frac{1}{4}y]\mathrm{\Psi }=0,$$
(13)
sau, definind :
$$\kappa =\frac{k^2}{2\lambda }=\frac{\overline{k}^2}{2m\omega }=\frac{E}{\mathrm{}\omega },$$
(14)
obลฃinem:
$$y\frac{d^2\mathrm{\Psi }}{dy^2}+\frac{1}{2}\frac{d\mathrm{\Psi }}{dy}+[\frac{\kappa }{2}\frac{1}{4}y]\mathrm{\Psi }=0.$$
(15)
Sฤ trecem la rezolvarea acestei ecuaลฃii, efectuรฎnd mai รฎntรฎi analiza sa asimptoticฤ รฎn limita $`y\mathrm{}`$. Pentru aceasta. se rescrie ecuaลฃia anterioarฤ รฎn forma :
$$\frac{d^2\mathrm{\Psi }}{dy^2}+\frac{1}{2y}\frac{d\mathrm{\Psi }}{dy}+[\frac{\kappa }{2y}\frac{1}{4}]\mathrm{\Psi }=0.$$
(16)
Observฤm cฤ รฎn limita $`y\mathrm{}`$ ecuaลฃia se comportฤ astfel:
$$\frac{d^2\mathrm{\Psi }_{\mathrm{}}}{dy^2}\frac{1}{4}\mathrm{\Psi }_{\mathrm{}}=0.$$
(17)
Aceastฤ ecuaลฃie are ca soluลฃie:
$$\mathrm{\Psi }_{\mathrm{}}(y)=A\mathrm{exp}\frac{y}{2}+B\mathrm{exp}\frac{y}{2}.$$
(18)
Eliminฤm $`\mathrm{exp}\frac{y}{2}`$ luรฎnd $`A=0`$ pentru cฤ diverge รฎn limita $`y\mathrm{}`$ ลi rฤmรฎnem cu exponenลฃiala atenuatฤ . Putem sugera acum cฤ $`\mathrm{\Psi }`$ are urmฤtoarea forma:
$$\mathrm{\Psi }(y)=\mathrm{exp}\frac{y}{2}\psi (y).$$
(19)
Substituind-o รฎn ecuaลฃia diferenลฃialฤ pentru $`y`$ ( ec. $`15`$) se obลฃine:
$$y\frac{d^2\psi }{dy^2}+(\frac{1}{2}y)\frac{d\psi }{dy}+(\frac{\kappa }{2}\frac{1}{4})\psi =0.$$
(20)
Ceea ce am obลฃinut este ecuaลฃia hipergeometricฤ confluentฤ <sup>4</sup><sup>4</sup>4Deasemenea cunoscutฤ ca ecuaลฃia diferenลฃialฤ Kummer. :
$$z\frac{d^2y}{dz^2}+(cz)\frac{dy}{dz}ay=0.$$
(21)
Soluลฃia generalฤ a acestei ecuaลฃii este :
$$y(z)=A_1F_1(a;c,z)+Bz_1^{1c}F_1(ac+1;2c,z),$$
(22)
unde funcลฃia hipergeometricฤ confluentฤ este definitฤ prin :
$${}_{1}{}^{}F_{1}^{}(a;c,z)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{(a)_nx^n}{(c)_nn!}.$$
(23)
Comparรฎnd acum ecuaลฃia noastrฤ, cu ecuaลฃia hipergeometricฤ confluentฤ, se observฤ cฤ soluลฃia generalฤ a primei este :
$$\psi (y)=A_1F_1(a;\frac{1}{2},y)+By_1^{\frac{1}{2}}F_1(a+\frac{1}{2};\frac{3}{2},y)$$
(24)
unde
$$a=(\frac{\kappa }{2}\frac{1}{4}).$$
(25)
Dacฤ menลฃinem aceste soluลฃii รฎn forma รฎn care se prezintฤ , condiลฃia de normalizare pentru funcลฃia de undฤ nu se satisface, pentru cฤ din comportamentul asimptotic al funcลฃiei hipergeometrice confluente <sup>5</sup><sup>5</sup>5 Comportamentul asimptotic pentru $`x\mathrm{}`$ este: $`{}_{1}{}^{}F_{1}^{}(a;c,z)\frac{\mathrm{\Gamma }(c)}{\mathrm{\Gamma }(ca)}e^{ia\pi }x^a+\frac{\mathrm{\Gamma }(c)}{\mathrm{\Gamma }(a)}e^xx^{ac}.`$ rezultฤ ( considerรฎnd numai comportamentul dominant exponenลฃial ) :
$$\mathrm{\Psi }(y)=e^{\frac{y}{2}}\psi (y)const.e^{\frac{y}{2}}y^{a\frac{1}{2}}.$$
(26)
Aceastฤ ultimฤ aproximaลฃie ne duce la o divergenลฃฤ รฎn integrala de normalizare, care fizic este inacceptabilฤ . Ceea ce se face รฎn acest caz, este sฤ se impunฤ condiลฃia de terminare a seriei <sup>6</sup><sup>6</sup>6Condiลฃia de truncare a seriei pentru funcลฃia hipergeometricฤ confluentฤ $`{}_{1}{}^{}F_{1}^{}(a;c,z)`$ este $`a=n`$, cu $`n`$ un รฎntreg nenegativ ( adicฤ , include zero ). , adicฤ , seria are numai un numฤr finit de termeni fiind deci un polinom de grad $`n`$.
Observฤm astfel cฤ faptul de a cere ca integrala de normalizare sฤ fie finitฤ (dupฤ cum ลtim condiลฃie obligatorie pentru semnificaลฃia fizicฤ รฎn termeni de probabilitฤลฃi), ne conduce la truncarea seriei, fapt care รฎn acelaลi timp produce cuantizarea energiei.
Considerฤm รฎn continuare cele douฤ cazuri posibile :
$`1)a=n`$ ลi $`B=0`$
$$\frac{\kappa }{2}\frac{1}{4}=n.$$
(27)
FP-urile sunt date de:
$$\mathrm{\Psi }_n(x)=D_n\mathrm{exp}\frac{\lambda x^2}{2}_1F_1(n;\frac{1}{2},\lambda x^2)$$
(28)
ลi energia este:
$$E_n=\mathrm{}\omega (2n+\frac{1}{2}).$$
(29)
$`2)a+\frac{1}{2}=n`$ ลi $`A=0`$
$$\frac{\kappa }{2}\frac{1}{4}=n+\frac{1}{2}.$$
(30)
FP-urile sunt acum:
$$\mathrm{\Psi }_n(x)=D_n\mathrm{exp}\frac{\lambda x^2}{2}x_1F_1(n;\frac{3}{2},\lambda x^2),$$
(31)
iar energiile staลฃionare sunt:
$$E_n=\mathrm{}\omega [(2n+1)+\frac{1}{2}].$$
(32)
Polinoamele obลฃinute รฎn urma acestei truncฤri a seriei hipergeometrice se numesc polinoame Hermite ลi se pot scrie ca urmฤtoarele funcลฃii hipergeometrice :
$$H_{2n}(\eta )=(1)^n\frac{(2n)!}{n!}_1F_1(n;\frac{1}{2},\eta ^2)$$
(33)
$$H_{2n1}(\eta )=(1)^n\frac{2(2n+1)!}{n!}\eta _1F_1(n;\frac{3}{2},\eta ^2).$$
(34)
Putem acum combina rezultatele obลฃinute ( pentru cฤ unele ne dau valorile pare ลi altele pe cele impare ) รฎntr-o singurฤ expresie pentru autovalori ลi funcลฃii proprii :
$$\mathrm{\Psi }_n(x)=D_n\mathrm{exp}\frac{\lambda x^2}{2}H_n(\sqrt{\lambda }x)$$
(35)
$$E_n=(n+\frac{1}{2})\mathrm{}\omega n=0,1,2\mathrm{}$$
(36)
Spectrul de energie al OA este echidistant, adicฤ , existฤ aceeaลi diferenลฃฤ $`\mathrm{}\omega `$ รฎntre oricare douฤ nivele. Altฤ observaลฃie pe care o putem face, este รฎn legฤturฤ cu valoarea minimฤ de energie pe care o are oscilatorul; poate รฎn mod surprinzฤtor este diferitฤ de zero; acesta se considerฤ un rezultat pur cuantic, pentru cฤ dispare dacฤ $`\mathrm{}0`$. Se cunoaลte ca energia de punct zero ลi faptul cฤ este diferitฤ de zero , este o caracteristicฤ a tuturor potenลฃialelor confinante .
Constanta de normalizare poate fi calculatฤ uลor ลi are valoarea:
$$D_n=\left[\sqrt{\frac{\lambda }{\pi }}\frac{1}{2^nn!}\right]^{\frac{1}{2}}.$$
(37)
Prin urmare se obลฃin funcลฃiile proprii normalizate ale OA unidimensional :
$$\mathrm{\Psi }_n(x)=\left[\sqrt{\frac{\lambda }{\pi }}\frac{1}{2^nn!}\right]^{\frac{1}{2}}\mathrm{exp}(\frac{\lambda x^2}{2})H_n(\sqrt{\lambda }x).$$
(38)
## Operatori de creare $`\widehat{a}^{}`$ ลi anihilare $`\widehat{a}`$
Existฤ o altฤ formฤ de a trata oscilatorul armonic faลฃฤ de cea convenลฃionalฤ de a rezolva ecuaลฃia Schrรถdinger. Este vorba de metoda algebricฤ sau metoda operatorilor de scarฤ , o metodฤ foarte eficientฤ care se poate aplica cu mult succes a multe probleme de mecanicฤ cuanticฤ de spectru discret.
Definim doi operatori nehermitici $`a`$ ลi $`a^{}`$ :
$$a=\sqrt{\frac{m\omega }{2\mathrm{}}}(x+\frac{ip}{m\omega })$$
(39)
$$a^{}=\sqrt{\frac{m\omega }{2\mathrm{}}}(x\frac{ip}{m\omega }).$$
(40)
Aceลti operatori sunt cunoscuลฃi ca operator de anihilare ลi operator de creare, respectiv (justificarea acestor denumiri se va vedea mai departe, deลi se poate spune cฤ vine din teoria cuanticฤ a cรฎmpurilor ).
Sฤ calculฤm acum comutatorul acestor doi operatori:
$$[a,a^{}]=\frac{m\omega }{2\mathrm{}}[x+\frac{ip}{m\omega },x\frac{ip}{m\omega }]=\frac{1}{2\mathrm{}}(i[x,p]+i[p,x])=1,$$
(41)
unde am folosit comutatorul:
$$[x,p]=i\mathrm{}.$$
(42)
Prin urmare operatorii de creare ลi anihilare nu comutฤ , satisfฤcรฎnd relaลฃiile de comutare :
$$[a,a^{}]=1.$$
(43)
Sฤ definim deasemenea importantul operator de numฤr $`\widehat{N}`$:
$$\widehat{N}=a^{}a.$$
(44)
Acest operator este hermitic dupฤ cum se poate demonstra uลor folosind $`(AB)^{}=B^{}A^{}`$ :
$$\widehat{N}^{}=(a^{}a)^{}=a^{}(a^{})^{}=a^{}a=\widehat{N}.$$
(45)
Considerรฎnd acum cฤ :
$$a^{}a=\frac{m\omega }{2\mathrm{}}(x^2+\frac{p^2}{m^2\omega ^2})+\frac{i}{2\mathrm{}}[x,p]=\frac{\widehat{H}}{\mathrm{}\omega }\frac{1}{2}$$
(46)
observฤm cฤ Hamiltonianul se scrie รฎntr-o formฤ simplฤ รฎn funcลฃie de operatorul de numฤr :
$$\widehat{H}=\mathrm{}\omega (\widehat{N}+\frac{1}{2}).$$
(47)
Operatorul de numฤr are acest nume datoritฤ faptului cฤ autovalorile sale sunt exact subindicii funcลฃiei de undฤ asupra cฤreia acลฃioneazฤ :
$$\widehat{N}n>=nn>,$$
(48)
unde am folosit notaลฃia:
$$\mathrm{\Psi }_n>=n>.$$
(49)
Aplicรฎnd acest fapt lui $`(47)`$ avem :
$$\widehat{H}n>=\mathrm{}\omega (n+\frac{1}{2})n>.$$
(50)
Dar ลtim din ecuaลฃia Schrรถdinger cฤ $`\widehat{H}n>=En>`$ pe baza cฤreia rezultฤ cฤ autovalorile energetice sunt date de :
$$E_n=\mathrm{}\omega (n+\frac{1}{2}).$$
(51)
Acest rezultat este identic (cum ลi trebuia sฤ fie ) cu rezultatul $`(36)`$.
รn continuare sฤ arฤtฤm de ce operatorii $`a`$ ลi $`a^{}`$ au numele pe care le au. Pentru aceasta sฤ calculฤm comutatorii:
$$[\widehat{N},a]=[a^{}a,a]=a^{}[a,a]+[a^{},a]a=a,$$
(52)
rezultate care se obลฃin din $`[a,a]=0`$ ลi $`(43)`$. Similar, sฤ calculฤm:
$$[\widehat{N},a^{}]=[a^{}a,a^{}]=a^{}[a,a^{}]+[a^{},a^{}]a=a^{}.$$
(53)
Cu aceลti doi comutatori putem sฤ scriem:
$`\widehat{N}(a^{}n>)`$ $`=`$ $`([\widehat{N},a^{}]+a^{}\widehat{N})n>`$
$`=`$ $`(a^{}+a^{}\widehat{N})n>`$
$`=`$ $`a^{}(1+n)n>=(n+1)a^{}n>.`$
Cu un procedeu similar se obลฃine deasemenea:
$$\widehat{N}(an>)=([\widehat{N},a]+a\widehat{N})n>=(n1)an>.$$
(55)
Expresia $`(54)`$ implicฤ cฤ se poate considera ket-ul $`a^{}n>`$ ca eigenket al operatorului de numฤr , unde autovaloarea incrementฤ cu unu, adicฤ , a fost produsฤ o cuantฤ de energie prin acลฃiunea lui $`a^{}`$ asupra ket-ului. Aceasta explicฤ numele de operator de creare (creaลฃie). Comentarii urmรฎnd aceeaลi linie de raลฃionament acelaลi tip de concluzie rezultฤ pentru operatorul $`a`$, ceea ce รฎi dฤ numele de operator de anihilare ( o cuantฤ de energie este eliminatฤ cรฎnd acลฃioneazฤ acest operator ).
Ecuaลฃia $`(54)`$ deasemenea implicฤ proporลฃionalitatea ket-urilor $`a^{}n>`$ ลi $`n+1>`$:
$$a^{}n>=cn+1>,$$
(56)
unde $`c`$ este o constantฤ care trebuie determinatฤ . Considerรฎnd รฎn plus cฤ :
$$(a^{}n>)^{}=<na=c^{}<n+1,$$
(57)
putem realiza urmฤtorul calcul:
$$<na(a^{}n>)=c^{}<n+1(cn+1>)$$
(58)
$$<naa^{}n>=c^{}c<n+1n+1>$$
(59)
$$<naa^{}n>=c^2.$$
(60)
Dar din relaลฃia de comutare pentru operatorii $`a`$ ลi $`a^{}`$ :
$$[a,a^{}]=aa^{}a^{}a=aa^{}\widehat{N}=1,$$
(61)
avem cฤ :
$$aa^{}=\widehat{N}+1$$
(62)
Substituind รฎn $`(60)`$:
$$<n\widehat{N}+1n>=<nn>+<n\widehat{N}n>=n+1=c^2.$$
(63)
Cerรฎnd $`c`$ sฤ fie real ลi pozitiv ( prin convenลฃie ), obลฃinem urmฤtoarea valoare:
$$c=\sqrt{n+1}.$$
(64)
Cu aceasta avem relaลฃia:
$$a^{}n>=\sqrt{n+1}n+1>.$$
(65)
Urmรฎnd acelaลi procedeu se poate ajunge la o relaลฃie pentru operatorul de anihilare :
$$an>=\sqrt{n}n1>.$$
(66)
Sฤ arฤtฤm acum cฤ valorile lui $`n`$ trebuie sฤ fie รฎntregi nenegativi. Pentru aceasta, recurgem la cerinลฃa de pozitivitate a normei, aplicรฎnd-o รฎn special vectorului de stare $`an>`$. Aceastฤ condiลฃie ne spune cฤ produsul interior (intern) al acestui vector cu adjunctul sฤu ($`(an>)^{}=<na^{}`$) trebuie sฤ fie mai mare sau egalฤ cu zero :
$$(<na^{})(an>)0.$$
(67)
Dar aceastฤ relaลฃie nu este decรฎt :
$$<na^{}an>=<n\widehat{N}n>=n0.$$
(68)
Prin urmare $`n`$ nu poate fi negativ ลi trebuie sฤ fie รฎntreg pentru cฤ dacฤ nu ar fi prin aplicarea consecutivฤ a operatorului de anihilare ne-ar duce la valori negative ale lui $`n`$, ceea ce este รฎn contradicลฃie cu ce s-a spus anterior.
Este posibil sฤ se exprime starea $`n`$ $`(n>)`$ direct รฎn funcลฃie de starea bazฤ $`(0>)`$ folosind operatorul de creare. Sฤ vedem cum se face aceastฤ iteraลฃie importantฤ :
$`1>=a^{}0>`$ (69)
$`2>=[{\displaystyle \frac{a^{}}{\sqrt{2}}}]1>=[{\displaystyle \frac{(a^{})^2}{\sqrt{2!}}}]0>`$ (70)
$`3>=[{\displaystyle \frac{a^{}}{\sqrt{3}}}]2>=[{\displaystyle \frac{(a^{})^3}{\sqrt{3!}}}]0>`$ (71)
$`n>=[{\displaystyle \frac{(a^{})^n}{\sqrt{n!}}}]0>`$ (72)
Putem deasemenea aplica aceastฤ metodฤ pentru a obลฃine FP-urile รฎn spaลฃiul configuraลฃiilor. Pentru a realiza acest lucru, vom pleca din starea bazฤ :
$$a0>=0.$$
(73)
รn reprezentarea $`x`$ avem:
$$\widehat{a}\mathrm{\Psi }_0(x)=\sqrt{\frac{m\omega }{2\mathrm{}}}(x+\frac{ip}{m\omega })\mathrm{\Psi }_0(x)=0.$$
(74)
Amintindu-ne forma pe care o ia operatorul impuls รฎn reprezentarea $`x`$, putem ajunge la o ecuaลฃie diferenลฃialฤ pentru funcลฃia de undฤ a stฤrii fundamentale; sฤ introducem deasemenea urmฤtoarea definiลฃie $`x_0=\sqrt{\frac{\mathrm{}}{m\omega }}`$ , cu care avem :
$$(x+x_0^2\frac{d}{dx})\mathrm{\Psi }_0=0$$
(75)
Aceastฤ ecuaลฃie se poate rezolva uลor, ลi prin normalizare ( integrala sa de la $`\mathrm{}`$ la $`\mathrm{}`$ trebuie sฤ fie pusฤ egalฤ cu unu ), ajungem la funcลฃia de undฤ a stฤrii fundamentale :
$$\mathrm{\Psi }_0(x)=(\frac{1}{\sqrt{\sqrt{\pi }x_0}})e^{\frac{1}{2}(\frac{x}{x_0})^2}$$
(76)
Restul de FP, care descriu stฤrile excitate ale OA , se pot obลฃine folosind operatorul de creaลฃie. Procedeul este urmฤtorul:
$`\mathrm{\Psi }_1=a^{}\mathrm{\Psi }_0=({\displaystyle \frac{1}{\sqrt{2}x_0}})(xx_0^2{\displaystyle \frac{d}{dx}})\mathrm{\Psi }_0`$ (77)
$`\mathrm{\Psi }_2={\displaystyle \frac{1}{\sqrt{2}}}(a^{})^2\mathrm{\Psi }_0={\displaystyle \frac{1}{\sqrt{2!}}}({\displaystyle \frac{1}{\sqrt{2}x_0}})^2(xx_0^2{\displaystyle \frac{d}{dx}})^2\mathrm{\Psi }_0.`$ (78)
Continuรฎnd, se poate arฤta prin inducลฃie cฤ :
$$\mathrm{\Psi }_n=\frac{1}{\sqrt{\sqrt{\pi }2^nn!}}\frac{1}{x_0^{n+\frac{1}{2}}}(xx_0^2\frac{d}{dx})^ne^{\frac{1}{2}(\frac{x}{x_0})^2}.$$
(79)
## Evoluลฃia temporalฤ a oscilatorului
รn aceastฤ secลฃiune vom ilustra prin intermediul OA modul รฎn care se lucreazฤ cu reprezentarea Heisenberg รฎn care stฤrile sunt fixate รฎn timp ลi se permite evoluลฃia temporalฤ a operatorilor. Vom considera operatorii ca funcลฃii de timp ลi vom obลฃine รฎn mod concret cum evoluลฃioneazฤ operatorii de poziลฃie , impuls, $`a`$ ลi $`a^{}`$ รฎn timp pentru cazul OA. Ecuaลฃiile de miลcare Heisenberg pentru $`p`$ ลi $`x`$ sunt :
$`{\displaystyle \frac{d\widehat{p}}{dt}}`$ $`=`$ $`{\displaystyle \frac{}{\widehat{x}}}V(\widehat{๐ฑ})`$ (80)
$`{\displaystyle \frac{d\widehat{x}}{dt}}`$ $`=`$ $`{\displaystyle \frac{\widehat{p}}{m}}.`$ (81)
De aici rezultฤ cฤ ecuaลฃiile de miลcare pentru $`x`$ y $`p`$ รฎn cazul OA sunt:
$`{\displaystyle \frac{d\widehat{p}}{dt}}`$ $`=`$ $`m\omega ^2\widehat{x}`$ (82)
$`{\displaystyle \frac{d\widehat{x}}{dt}}`$ $`=`$ $`{\displaystyle \frac{\widehat{p}}{m}}.`$ (83)
Prin urmare, dispunem de o pereche de ecuaลฃii cuplate , care sunt echivalente unei perechi de ecuaลฃii pentru operatorii de creaลฃie ลi anihilare, care รฎnsฤ nu sunt cuplate. รn mod explicit :
$`{\displaystyle \frac{da}{dt}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{m\omega }{2\mathrm{}}}}{\displaystyle \frac{d}{dt}}(\widehat{x}+{\displaystyle \frac{i\widehat{p}}{m\omega }})`$ (84)
$`{\displaystyle \frac{da}{dt}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{m\omega }{2\mathrm{}}}}({\displaystyle \frac{d\widehat{x}}{dt}}+{\displaystyle \frac{i}{m\omega }}{\displaystyle \frac{d\widehat{p}}{dt}}).`$ (85)
Substituind $`(82)`$ ลi $`(83)`$ รฎn $`(85)`$ :
$$\frac{da}{dt}=\sqrt{\frac{m\omega }{2\mathrm{}}}(\frac{\widehat{p}}{m}i\omega \widehat{x})=i\omega a.$$
(86)
Similar, se poate obลฃine o ecuaลฃie diferenลฃialฤ pentru operatorul de creaลฃie :
$$\frac{da^{}}{dt}=i\omega a^{}$$
(87)
Ecuaลฃiile diferenลฃiale pe care le-am obลฃinut pentru evoluลฃia temporalฤ a operatorilor de creaลฃie ลi anihilare , pot fi integrate imediat, dรฎnd evoluลฃia explicitฤ a acestor operatori:
$`a(t)`$ $`=`$ $`a(0)e^{i\omega t}`$ (88)
$`a^{}(t)`$ $`=`$ $`a^{}(0)e^{i\omega t}.`$ (89)
Se poate remarca din aceste rezultate ลi din ecuaลฃiile $`(44)`$ ลi $`(47)`$ cฤ atรฎt Hamiltonianul ca ลi operatorul de numฤr , nu depind de timp, aลa cum era de aลteptat.
Cu cele douฤ rezultate anterioare , putem sฤ obลฃinem operatorii de poziลฃie ลi impuls ca funcลฃii de timp, pentru cฤ sunt daลฃi รฎn funcลฃie de operatorii de creaลฃie ลi anihilare:
$`\widehat{x}`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mathrm{}}{2m\omega }}}(a+a^{})`$ (90)
$`\widehat{p}`$ $`=`$ $`i\sqrt{{\displaystyle \frac{m\mathrm{}\omega }{2}}}(a^{}a).`$ (91)
Substituind-i se obลฃine:
$`\widehat{x}(t)`$ $`=`$ $`\widehat{x}(0)\mathrm{cos}\omega t+{\displaystyle \frac{\widehat{p}(0)}{m\omega }}\mathrm{sin}\omega t`$ (92)
$`\widehat{p}(t)`$ $`=`$ $`m\omega \widehat{x}(0)\mathrm{sin}\omega t+\widehat{p}(0)\mathrm{cos}\omega t.`$ (93)
Evoluลฃia temporalฤ a acestor operatori este aceeaลi ca รฎn cazul ecuaลฃiilor clasice de miลcare.
Astfel, am arฤtat forma explicitฤ de evoluลฃie a patru operatori bazici รฎn cazul OA , arฤtรฎnd modul รฎn care se lucreazฤ รฎn reprezentarea Heisenberg.
## OA tridimensional
La รฎnceputul analizei noastre a OA cuantic am fฤcut comentarii รฎn legฤturฤ cu importanลฃa pentru fizicฤ a OA . Dacฤ vom considera un analog tridimensional, ar trebui sฤ considerฤm o dezvoltare Taylor รฎn trei variabile<sup>7</sup><sup>7</sup>7Este posibil sฤ se exprime dezvoltarea Taylor รฎn jurul lui $`๐ซ_\mathrm{๐}`$ ca un operator exponenลฃial : $`e^{[(xx_o)+(yy_o)+(zz_o)](\frac{}{x}+\frac{}{y}+\frac{}{z})}f(๐ซ_๐จ).`$ reลฃinรฎnd termeni numai pรฎnฤ รฎn ordinul doi inclusiv, ceea ce obลฃinem este o formฤ cuadraticฤ (รฎn cazul cel mai general). Problema de rezolvat รฎn aceastฤ aproximaลฃie nu este chiar atรฎt de simplฤ cum ar pฤrea dintr-o primฤ examinare a potenลฃialului corespunzฤtor :
$$V(x,y,z)=ax^2+by^2+cz^2+dxy+exz+fyz.$$
(94)
Existฤ รฎnsฤ multe sisteme care posedฤ simetrie sfericฤ sau pentru care aproximaลฃia acestei simetrii este satisfฤcฤtoare. รn acest caz:
$$V(x,y,z)=K(x^2+y^2+z^2),$$
(95)
ceea ce este echivalent cu a spune cฤ derivatele parลฃiale secunde ( nemixte ) iau toate aceeaลi valoare ( รฎn cazul anterior reprezentate prin $`K`$). Putem adฤuga cฤ aceastฤ este o bunฤ aproximaลฃie รฎn cazul รฎn care valorile derivatelor parลฃiale secunde mixte sunt mici รฎn comparaลฃie cu cele nemixte.
Cรฎnd se satisfac aceste condiลฃii ลi potenลฃialul este dat de $`(95)`$ spunem cฤ sistemul este un OA tridimensional sferic simetric.
Hamiltonianul pentru acest caz este de forma:
$$\widehat{H}=\frac{\mathrm{}^2}{2m}^2+\frac{m\omega ^2}{2}r^2,$$
(96)
unde Laplaceanul este dat รฎn coordonate sferice ลi $`r`$ este variabila sfericฤ radialฤ .
Fiind vorba de un potenลฃial independent de timp, energia se conservฤ ; รฎn plus datฤ simetria sfericฤ , momentul cinetic deasemenea se conservฤ . Avem deci douฤ mฤrimi conservate, ceea ce ne permite sฤ spunem cฤ fiecฤreia รฎi corespunde un numฤr cuantic. Putem sฤ presupunem cฤ funcลฃiile de undฤ depind de douฤ numere cuantice (deลi รฎn acest caz vom vedea cฤ apare รฎncฤ unul ). Cu aceste comentarii, ecuaลฃia de interes este :
$$\widehat{H}\mathrm{\Psi }_{nl}=E_{nl}\mathrm{\Psi }_{nl}.$$
(97)
Laplaceanul รฎn coordonate sferice este :
$$^2=\frac{^2}{r^2}+\frac{2}{r}\frac{}{r}\frac{\widehat{L}^2}{\mathrm{}^2r^2}$$
(98)
ลi rezultฤ din faptul cunoscut :
$$\widehat{L}^2=\mathrm{}^2[\frac{1}{\mathrm{sin}\theta }\frac{}{\theta }(\mathrm{sin}\theta \frac{}{\theta })+\frac{1}{\mathrm{sin}\theta ^2}\frac{^2}{\phi ^2}].$$
(99)
Funcลฃiile proprii ale lui $`\widehat{L}^2`$ sunt armonicele sferice, respectiv:
$$\widehat{L}^2Y_{lm_l}(\theta ,\phi )=\mathrm{}^2l(l+1)Y_{lm_l}(\theta ,\phi )$$
(100)
Faptul cฤ armonicele sferice poartฤ numฤrul cuantic $`m_l`$ face ca acesta sฤ fie introdus รฎn funcลฃia de undฤ $`\mathrm{\Psi }_{nlm_l}`$.
Pentru a realiza separarea variabilelor ลi funcลฃiilor se propune substituลฃia:
$$\mathrm{\Psi }_{nlm_l}(r,\theta ,\phi )=\frac{R_{nl}(r)}{r}Y_{lm_l}(\theta ,\phi ).$$
(101)
Odatฤ introdusฤ รฎn ecuaลฃia Schrรถdinger va separa partea spaลฃialฤ de cea unghiularฤ ; ultima se identificฤ cu un operator proporลฃional cu operatorul moment cinetic pฤtrat, pentru care funcลฃiile proprii sunt armonicele sferice, รฎn timp ce รฎn partea spaลฃialฤ obลฃinem ecuaลฃia :
$$R_{nl}^{\prime \prime }+(\frac{2mE_{nl}}{\mathrm{}^2}\frac{m^2\omega ^2}{\mathrm{}^2}r^2\frac{l(l+1)}{r^2})R_{nl}(r)=0.$$
(102)
Folosind definiลฃiile $`(7)`$ ลi $`(8)`$ , ecuaลฃia anterioarฤ ia exact forma lui $`(9)`$, cu excepลฃia termenului de moment unghiular, care รฎn mod comun se cunoaลte ca barierฤ de moment unghiular.
$$R_{nl}^{\prime \prime }+(k^2\lambda ^2r^2\frac{l(l+1)}{r^2})R_{nl}=0.$$
(103)
Pentru a rezolva aceastฤ ecuaลฃie , vom pleca dela analiza sa asimptoticฤ . Dacฤ vom considera mai รฎntรฎi $`r\mathrm{}`$, observฤm cฤ termenul de moment unghiular este neglijabil, astfel cฤ รฎn aceastฤ limitฤ comportamentul asimptotic este identic aceluia a lui $`(9)`$, ceea ce ne conduce la:
$$R_{nl}(r)\mathrm{exp}\frac{\lambda r^2}{2}\text{รฎn}limr\mathrm{}.$$
(104)
Dacฤ studiem acum comportamentul รฎn jurul lui zero, vedem cฤ termenul dominant este cel de moment unghiular, adicฤ , ecuaลฃia diferenลฃialฤ $`(102)`$ se converte รฎn aceastฤ limitฤ รฎn :
$$R_{nl}^{\prime \prime }\frac{l(l+1)}{r^2}R_{nl}=0.$$
(105)
Aceasta este o ecuaลฃie diferenลฃialฤ de tip Euler <sup>8</sup><sup>8</sup>8O ecuaลฃie de tip Euler este de forma :
$$x^ny^{(n)}(x)+x^{n1}y^{(n1)}(x)+\mathrm{}+xy^{}(x)+y(x)=0.$$
Soluลฃiile ei sunt de tipul $`x^\alpha `$, care se substituie ลi se gฤseลte un polinom รฎn $`\alpha `$. , a cฤrei rezolvare duce la douฤ soluลฃii independente:
$$R_{nl}(r)r^{l+1}\text{sau}r^l\text{รฎn}limr0.$$
(106)
Argumentele anterioare ne conduc la a propune substituลฃia :
$$R_{nl}(r)=r^{l+1}\mathrm{exp}\frac{\lambda r^2}{2}\varphi (r).$$
(107)
S-ar putea deasemenea face ลi urmฤtoarea substituลฃie:
$$R_{nl}(r)=r^l\mathrm{exp}\frac{\lambda r^2}{2}v(r),$$
(108)
care รฎnsฤ ne conduce la aceleaลi soluลฃii ca ลi $`(107)`$ ( de arฤtat acest lucru este un bun exerciลฃiu). Substituind $`(107)`$ รฎn $`(103)`$ , se obลฃine urmฤtoarea ecuaลฃie diferenลฃialฤ pentru $`\varphi `$ :
$$\varphi ^{\prime \prime }+2(\frac{l+1}{r}\lambda r)\varphi ^{}[\lambda (2l+3)k^2]\varphi =0.$$
(109)
Cu schimbarea de variabilฤ $`w=\lambda r^2`$, obลฃinem:
$$w\varphi ^{\prime \prime }+(l+\frac{3}{2}w)\varphi ^{}[\frac{1}{2}(l+\frac{3}{2})\frac{\kappa }{2}]\varphi =0,$$
(110)
unde am introdus $`\kappa =\frac{k^2}{2\lambda }=\frac{E}{\mathrm{}\omega }`$. Am ajuns din nou la o ecuaลฃie diferenลฃialฤ de tip hipergeometric confluentฤ cu soluลฃiile ( a se vedea $`(21)`$ ลi $`(22)`$):
$$\varphi (r)=A_1F_1[\frac{1}{2}(l+\frac{3}{2}\kappa );l+\frac{3}{2},\lambda r^2]+Br_1^{(2l+1)}F_1[\frac{1}{2}(l+\frac{1}{2}\kappa );l+\frac{1}{2},\lambda r^2].$$
(111)
A doua soluลฃie particularฤ nu poate fi normalizatฤ , pentru cฤ diverge puternic รฎn zero, ceea ce obligฤ a lua $`B=0`$, deci :
$$\varphi (r)=A_1F_1[\frac{1}{2}(l+\frac{3}{2}\kappa );l+\frac{3}{2},\lambda r^2].$$
(112)
Folosind aceleaลi argumente ca รฎn cazul OA unidimensional, respectiv, a impune ca soluลฃiile sฤ fie regulare รฎn infinit, รฎnseamnฤ condiลฃia de truncare a seriei, ceea ce implicฤ din nou cuantizarea energiei . Truncarea este รฎn acest caz:
$$\frac{1}{2}(l+\frac{3}{2}\kappa )=n,$$
(113)
unde introducรฎnd explicit $`\kappa `$, obลฃinem spectrul de energie :
$$E_{nl}=\mathrm{}\omega (2n+l+\frac{3}{2}).$$
(114)
Putem observa cฤ pentru OA tridimensional sferic simetric existฤ o energie de punct zero $`\frac{3}{2}\mathrm{}\omega `$.
Funcลฃiile proprii nenormalizate sunt:
$$\mathrm{\Psi }_{nlm}(r,\theta ,\phi )=r^le_1^{\frac{\lambda r^2}{2}}F_1(n;l+\frac{3}{2},\lambda r^2)Y_{lm}(\theta ,\phi ).$$
(115)
## 5P. Probleme
### Problema 5.1
Sฤ se determine autovalorile ลi funcลฃiile proprii ale OA รฎn spaลฃiul impulsurilor.
Hamiltonianul cuantic de OA este:
$$\widehat{H}=\frac{\widehat{p}^2}{2m}+\frac{1}{2}m\omega ^2\widehat{x}^2.$$
รn spaลฃiul impulsurilor, operatorii $`\widehat{x}`$ ลi $`\widehat{p}`$ au urmฤtoarea formฤ :
$$\widehat{p}p$$
$$\widehat{x}i\mathrm{}\frac{}{p}.$$
Prin urmare Hamiltonianul cuantic OA รฎn reprezentarea de impuls este :
$$\widehat{H}=\frac{p^2}{2m}\frac{1}{2}m\omega ^2\mathrm{}^2\frac{d^2}{dp^2}.$$
Avem de rezolvat problema de autovalori ( ceea ce รฎnseamnฤ de obลฃinut funcลฃiile proprii ลi autovalorile) datฤ prin $`(5)`$ , care cu Hamiltonianul anterior, este urmฤtoarea ecuaลฃie diferenลฃialฤ :
$$\frac{d^2\mathrm{\Psi }(p)}{dp^2}+(\frac{2E}{m\mathrm{}^2\omega ^2}\frac{p^2}{m^2\mathrm{}^2\omega ^2})\mathrm{\Psi }(p)=0.$$
(116)
Se poate observa cฤ aceastฤ ecuaลฃie diferenลฃialฤ , este identicฤ , pรฎnฤ la constante, cu ecuaลฃia diferenลฃialฤ din spaลฃiul configuraลฃiilor ( ec. $`(6)`$ ). Pentru a exemplifica o altฤ formฤ de a o rezolva, nu vom urma exact acelaลi drum.
Definim doi parametri, analogi celor din $`(7)`$ ลi $`(8)`$:
$$k^2=\frac{2E}{m\mathrm{}^2\omega ^2}\lambda =\frac{1}{m\mathrm{}\omega }.$$
(117)
Cu aceste definiลฃii, ajungem la ecuaลฃia diferenลฃialฤ $`(9)`$ลi deci soluลฃia cฤutatฤ ( รฎn urma efectuฤrii analizei asimptotice ) este de forma:
$$\mathrm{\Psi }(y)=e^{\frac{1}{2}y}\varphi (y),$$
(118)
unde $`y`$ este dat de $`y=\lambda p^2`$ ลi $`\lambda `$ este definit รฎn $`(117)`$. Substituind $`(118)`$ รฎn $`(116)`$ , avรฎnd grijฤ sฤ punem $`(118)`$ รฎn variabila $`p`$ . Se obลฃine astfel o ecuaลฃie diferenลฃialฤ pentru $`\varphi `$ :
$$\frac{d^2\varphi (p)}{dp^2}2\lambda p\frac{d\varphi (p)}{dp}+(k^2\lambda )\varphi (p)=0.$$
(119)
Vom face acum schimbul de variabilฤ $`u=\sqrt{\lambda }p`$ , care ne conduce la ecuaลฃia Hermite :
$$\frac{d^2\varphi (u)}{du^2}2u\frac{d\varphi (u)}{du}+2n\varphi (u)=0,$$
(120)
cu $`n`$ un รฎntreg nenegativ , ลi unde am pus :
$$\frac{k^2}{\lambda }1=2n.$$
De aici ลi din definiลฃiile date รฎn $`(117)`$ rezultฤ cฤ autovalorile de energie sunt date de :
$$E_n=\mathrm{}\omega (n+\frac{1}{2}).$$
Soluลฃiile pentru $`(120)`$ sunt polinoamele Hermite $`\varphi (u)=H_n(u)`$ ลi funcลฃiile proprii nenormalizate sunt :
$$\mathrm{\Psi }(p)=Ae^{\frac{\lambda }{2}p^2}H_n(\sqrt{\lambda }p).$$
### Problema 5.2
Sฤ se demonstreze cฤ polinoamele Hermite pot fi expresate รฎn urmฤtoarea reprezentare integralฤ :
$$H_n(x)=\frac{2^n}{\sqrt{\pi }}_{\mathrm{}}^{\mathrm{}}(x+iy)^ne^{y^2}๐y.$$
(121)
Aceastฤ reprezentare a polinoamelor Hermite nu este foarte uzualฤ , deลi se poate dovedi utilฤ รฎn unele cazuri. Ceea ce vom face pentru a realiza demonstraลฃia , este sฤ dezvoltฤm integrala ลi sฤ demonstrฤm cฤ ceea ce se obลฃine este identic cu reprezentarea รฎn serie a polinoamelor Hermite pentru care avem :
$$\underset{k=0}{\overset{[\frac{n}{2}]}{}}\frac{(1)^kn!}{(n2k)!k!}(2x)^{n2k},$$
(122)
unde simbolul $`[c]`$ unde se terminฤ seria significฤ cel mai mare รฎntreg mai mic sau egal cu $`c`$.
Primul lucru pe care รฎl vom face este sฤ dezvoltฤm binomul din integralฤ folosind binecunoscuta teoremฤ a binomului:
$$(x+y)^n=\underset{m=0}{\overset{n}{}}\frac{n!}{(nm)!m!}x^{nm}y^m.$$
Astfel:
$$(x+iy)^n=\underset{m=0}{\overset{n}{}}\frac{n!}{(nm)!m!}i^mx^{nm}y^m,$$
(123)
care substituit รฎn integralฤ duce la:
$$\frac{2^n}{\sqrt{\pi }}\underset{m=0}{\overset{n}{}}\frac{n!}{(nm)!m!}i^mx^{nm}_{\mathrm{}}^{\mathrm{}}y^me^{y^2}๐y.$$
(124)
Din forma expresiei din integralฤ putem sฤ vedem cฤ este diferitฤ de zero cรฎnd $`m`$ este par , รฎn cazul impar integrala se anuleazฤ din motive evidente. Folosind notaลฃia parฤ $`m=2k`$ avem:
$$\frac{2^n}{\sqrt{\pi }}\underset{k=0}{\overset{[\frac{n}{2}]}{}}\frac{n!}{(n2k)!(2k)!}i^{2k}x^{n2k}2_0^{\mathrm{}}y^{2k}e^{y^2}๐y.$$
(125)
Cu schimbul de variabilฤ $`u=y^2`$, integrala devine o funcลฃie gamma :
$$\frac{2^n}{\sqrt{\pi }}\underset{k=0}{\overset{[\frac{n}{2}]}{}}\frac{n!}{(n2k)!(2k)!}i^{2k}x^{n2k}_0^{\mathrm{}}u^{k\frac{1}{2}}e^u๐u,$$
(126)
respectiv $`\mathrm{\Gamma }(k+\frac{1}{2})`$ , care รฎn plus se poate exprima prin factoriali ( desigur pentru $`k`$ รฎntreg ) :
$$\mathrm{\Gamma }(k+\frac{1}{2})=\frac{(2k)!}{2^{2k}k!}\sqrt{\pi }.$$
Substituind aceastฤ expresie รฎn sumฤ ลi folosind faptul cฤ $`i^{2k}=(1)^k`$ se obลฃine
$$\underset{k=0}{\overset{[\frac{n}{2}]}{}}\frac{(1)^kn!}{(n2k)!k!}(2x)^{n2k},$$
(127)
care este identic cu $`(122)`$ , ceea ce completeazฤ demonstraลฃia.
### Problema 5.3
Sฤ se arate cฤ relaลฃia de incertitudine Heisenberg se satisface efectuรฎnd calculul cu funcลฃiile proprii ale OA .
Trebuie sฤ arฤtฤm cฤ pentru oricare $`\mathrm{\Psi }_n`$ se satisface:
$$<(\mathrm{\Delta }p)^2(\mathrm{\Delta }x)^2>\frac{\mathrm{}^2}{4},$$
(128)
unde notaลฃia $`<>`$ รฎnseamnฤ valoare medie.
Vom calcula รฎn mod separat $`<(\mathrm{\Delta }p)^2>`$ ลi $`<(\mathrm{\Delta }x)^2>`$ , unde fiecare din aceste expresii este :
$$<(\mathrm{\Delta }p)^2>=<(p<p>)^2>=<p^22p<p>+<p>^2>=<p^2><p>^2,$$
$$<(\mathrm{\Delta }x)^2>=<(x<x>)^2>=<x^22x<x>+<x>^2>=<x^2><x>^2.$$
Mai รฎntรฎi vom arฤta cฤ atรฎt media lui $`x`$ cรฎt ลi a lui $`p`$ sunt zero. Pentru media lui $`x`$ avem:
$$<x>=_{\mathrm{}}^{\mathrm{}}x[\mathrm{\Psi }_n(x)]^2๐x.$$
Aceastฤ integralฤ se anuleazฤ datoritฤ imparitฤลฃtii expresiei de integrat, care este manifestฤ . Rezultฤ deci cฤ :
$$<x>=0.$$
(129)
Aceleaลi argumente sunt corecte pentru media lui $`p`$ , dacฤ efectuฤm calculul รฎn spaลฃiul impulsurilor, respectiv cu ajutorul funcลฃiilor obลฃinute รฎn problema 1 . Este suficient sฤ vedem cฤ forma funลฃionalฤ este aceaลi (se schimbฤ doar simbolul). Deci:
$$<p>=0.$$
(130)
Sฤ calculฤm acum media lui $`x^2`$. Vom folosi teorema virialului <sup>9</sup><sup>9</sup>9 Amintim cฤ teorema virialului รฎn mecanica cuanticฤ afirmฤ cฤ :
$$2<T>=<๐ซV(๐ซ)>.$$
Pentru un potenลฃial de forma $`V=\lambda x^n`$ se satisface:
$$2<T>=n<V>,$$
unde $`T`$ reprezintฤ energia cineticฤ ลi V este energia potenลฃialฤ .. Observฤm mai รฎntรฎi cฤ :
$$<V>=\frac{1}{2}m\omega ^2<x^2>.$$
Prin urmare este posibilฤ relaลฃionarea mediei lui $`x^2`$ direct cu media potenลฃialului รฎn acest caz (ลi deci folosirea teoremei virialului).
$$<x^2>=\frac{2}{m\omega ^2}<V>.$$
(131)
Avem nevoie deasemenea de media energiei totale :
$$<H>=<T>+<V>,$$
pentru care din nou se poate folosi teorema virialului ( pentru $`n=2`$ ) :
$$<H>=2<V>.$$
(132)
Astfel, se obลฃine:
$$<x^2>=\frac{<H>}{m\omega ^2}=\frac{\mathrm{}\omega (n+\frac{1}{2})}{m\omega ^2}$$
(133)
$$<x^2>=\frac{\mathrm{}}{m\omega }(n+\frac{1}{2}).$$
(134)
Similar, media lui $`p^2`$ se poate calcula explicit:
$$<p^2>=2m<\frac{p^2}{2m}>=2m<T>=m<H>=m\mathrm{}\omega (n+\frac{1}{2}).$$
(135)
Cu $`(133)`$ ลi $`(135)`$ avem:
$$<(\mathrm{\Delta }p)^2(\mathrm{\Delta }x)^2>=(n+\frac{1}{2})^2\mathrm{}^2.$$
(136)
Pe baza acestui rezultat ajungem la concluzia cฤ รฎn stฤrile staลฃionare ale OA, care practic nu au fost folosite รฎn mod direct, relaลฃia de incertitudine Heisenberg se satisface ลi are valoarea minimฤ pentru starea fundamentalฤ $`n=0`$.
### Problema 5.4
Sฤ se obลฃinฤ elementele de matrice ale operatorilor $`a`$, $`a^{}`$, $`\widehat{x}`$ ลi $`\widehat{p}`$.
Sฤ gฤsim mai รฎntรฎi elementele de matrice pentru operatorii de creaลฃie ลi anihilare, care sunt de mult ajutor pentru a obลฃine elemente de matrice pentru restul operatorilor.
Vom folosi relaลฃiile $`(65)`$ ลi $`(66)`$, care duc la:
$$<man>=\sqrt{n}<mn1>=\sqrt{n}\delta _{m,n1}.$$
(137)
รn mod similar pentru operatorul de creaลฃie avem rezultatul:
$$<ma^{}n>=\sqrt{n+1}<mn+1>=\sqrt{n+1}\delta _{m,n+1}.$$
(138)
Sฤ trecem acum la calculul elementelor de matrice ale operatorului de poziลฃie. Pentru al efectua , sฤ exprimฤm operatorul de posiลฃie รฎn funcลฃie de operatorii de creaลฃie ลi anihilare. Folosind definiลฃiile $`(39)`$ ลi $`(40)`$ , se demonstreazฤ imediat cฤ operatorul de posiลฃie este dat de :
$$\widehat{x}=\sqrt{\frac{\mathrm{}}{2m\omega }}(a+a^{}).$$
(139)
Folosind acest rezultat, elementele de matrice ale operatorului $`\widehat{x}`$ pot fi calculate รฎn manierฤ imediatฤ :
$`<m\widehat{x}n>`$ $`=`$ $`<m\sqrt{{\displaystyle \frac{\mathrm{}}{2m\omega }}}(a+a^{})n>`$ (140)
$`=`$ $`\sqrt{{\displaystyle \frac{\mathrm{}}{2m\omega }}}[\sqrt{n}\delta _{m,n1}+\sqrt{n+1}\delta _{m,n+1}].`$
Urmรฎnd acelaลi procedeu putem calcula elementele de matrice ale operatorului impuls, considerรฎnd cฤ $`\widehat{p}`$ este dat รฎn funcลฃie de operatorii de creaลฃie ลi anihilare รฎn forma :
$$\widehat{p}=i\sqrt{\frac{m\mathrm{}\omega }{2}}(a^{}a),$$
(141)
ceea ce ne conduce la:
$$<m\widehat{p}n>=i\sqrt{\frac{m\mathrm{}\omega }{2}}[\sqrt{n+1}\delta _{m,n+1}\sqrt{n}\delta _{m,n1}].$$
(142)
Se poate vedea uลurinลฃa cu care se pot face calculele dacฤ se folosesc elementele de matrice ale operatorilor de creaลฃie ลi anihilare. Finalizฤm cu o observaลฃie รฎn legฤturฤ cu nediagonalitatea elementelor de matrice obลฃinute. Aceasta este de aลteptat datoritฤ faptului cฤ reprezentarea folositฤ este cea a operatorului de numฤr ลi nici unul dintre cei patru operatori nu comutฤ cu el.
### Problema 5.5
Sฤ se gฤseascฤ valorile medii ale lui $`\widehat{x}^2`$ ลi $`\widehat{p}^2`$ pentru OA unidimensional ลi sฤ se foloseascฤ acestea pentru calculul valorilor medii (de aลteptare) ale energiei cinetice ลi celei potenลฃiale. Sฤ se compare acest ultim rezultat cu teorema virialului.
Mai รฎntรฎi sฤ obลฃinem valoarea medie a lui $`\widehat{x}^2`$. Pentru aceasta recurgem la expresia $`(139)`$, care ne conduce la :
$$\widehat{x}^2=\frac{\mathrm{}}{2m\omega }(a^2+(a^{})^2+a^{}a+aa^{}).$$
(143)
Se aminteลte cฤ operatorii de creaลฃie ลi anihilare nu comutฤ รฎntre ei . Avรฎnd $`(143)`$ putem calcula valoarea medie a lui $`\widehat{x}^2`$ :
$`<\widehat{x}^2>`$ $`=`$ $`<n\widehat{x}^2n>`$ (144)
$`=`$ $`{\displaystyle \frac{\mathrm{}}{2m\omega }}[\sqrt{n(n1)}\delta _{n,n2}+\sqrt{(n+1)(n+2)}\delta _{n,n+2}`$
$`+`$ $`n\delta _{n,n}+(n+1)\delta _{n,n}],`$
ceea ce aratฤ cฤ :
$$<\widehat{x}^2>=<n\widehat{x}^2n>=\frac{\mathrm{}}{2m\omega }(2n+1).$$
(145)
Pentru calcularea valorii medii a lui $`\widehat{p}^2`$ folosim $`(141)`$ pentru a exprima acest operator รฎn funcลฃie de operatorii de creaลฃie ลi anihilare:
$$\widehat{p}^2=\frac{m\mathrm{}\omega }{2}(a^2+(a^{})^2aa^{}a^{}a),$$
(146)
ceea ce ne conduce la:
$$<\widehat{p}^2>=<n\widehat{p}^2n>=\frac{m\mathrm{}\omega }{2}(2n+1).$$
(147)
Ultimul rezultat ne dฤ practic media energiei cinetice :
$$<\widehat{T}>=<\frac{\widehat{p}^2}{2m}>=\frac{1}{2m}<\widehat{p}^2>=\frac{\mathrm{}\omega }{4}(2n+1).$$
(148)
Valoarea medie a energiei potenลฃiale este:
$$<\widehat{V}>=<\frac{1}{2}m\omega ^2\widehat{x}^2>=\frac{1}{2}m\omega ^2<\widehat{x}^2>=\frac{\mathrm{}\omega }{4}(2n+1),$$
(149)
unde s-a folosit $`(145)`$.
Observฤm cฤ aceste valori medii coincid pentru orice $`n`$, ceea ce este รฎn conformitate cu teorema virialului, care ne spune cฤ pentru un potenลฃial cuadratic ca cel de OA, valorile medii ale energiei cinetice ลi potenลฃiale trebuie sฤ coincidฤ ลi deci sฤ fie jumฤtate din valoarea medie a energiei totale.
6. ATOMUL DE HIDROGEN
## Introducere
รn acest capitol vom studia atomul de hidrogen, rezolvรฎnd ecuaลฃia Schrรถdinger independentฤ de timp cu un potenลฃial produs de douฤ particule รฎncฤrcate electric cum este cazul electronului ลi protonului, cu Laplaceanul รฎn coordonate sferice. Din punct de vedere matematic, se va folosi metoda separฤrii de variabile, dรฎnd o interpretare fizicฤ funcลฃiei de undฤ ca soluลฃie a ecuaลฃiei Schrรถdinger pentru acest caz important, odatฤ cu interpretarea numerelor cuantice ลi a densitฤลฃilor de probabilitate.
Scala spaลฃialฤ foarte micฤ a atomului de hidrogen intrฤ รฎn domeniul de aplicabilitate al mecanicii cuantice, pentru care fenomenele atomice au fost o arie de verificare ลi interpretare a rezultatelor รฎncฤ de la bun รฎnceput. Cum mecanica cuanticฤ dฤ , รฎntre altele, relaลฃii รฎntre mฤrimile observabile ลi cum principiul de incertitudine modificฤ radical definiลฃia teoreticฤ a unei โobservabileโ este important sฤ รฎnลฃelegem รฎn mod cรฎt mai clar noลฃiunea cuanticฤ de observabilฤ รฎn cรฎmpul atomic. De acord cu principiul de incertitudine, poziลฃia ลi impulsul unei particule nu se pot mฤsura simultan sub o anumitฤ precizie impusฤ de comutatorii cuantici. De fapt, mฤrimile asupra cฤrora mecanica cuanticฤ dฤ rezultate ลi pe care le relaลฃioneazฤ sunt รฎntotdeauna probabilitฤลฃi. รn loc de a afirma, de exemplu, cฤ raza orbitei electronului รฎntr-o stare fundamentalฤ a atomului de hidrogen este รฎntotdeauna $`5.3\times 10^{11}`$ m, mecanica cuanticฤ afirmฤ cฤ aceasta este doar raza medie; dacฤ efectuฤm un experiment adecuat, vom obลฃine exact ca รฎn cazul experimentelor cu detectori macroscopici pe probe macroscopice diferite valori aleatorii dar a cฤror medie va fi $`5.3\times 10^{11}`$ m. Aลadar, din punctul de vedere al erorilor experimentale nu existฤ nici o diferenลฃฤ faลฃฤ de fizica clasicฤ .
Dupฤ cum se ลtie, pentru calculul valorilor medii รฎn mecanica cuanticฤ este necesarฤ o funcลฃie de undฤ corespunzฤtoare $`\mathrm{\Psi }`$. Deลi $`\mathrm{\Psi }`$ nu are o interpretare fizicฤ directฤ , modulul pฤtrat $`\mathrm{\Psi }^2`$ calculat รฎntr-un punct arbitrar din spaลฃiu ลi la un moment dat este proporลฃional cu probabilitatea de a gฤsi particula รฎntr-o vecinฤtate infinitezimalฤ a acelui punct acel loc ลi la momentul dat. Scopul mecanicii cuantice este determinarea lui $`\mathrm{\Psi }`$ pentru o microparticulฤ รฎn diferite condiลฃii experimentale.
รnainte de a trece la calculul efectiv al lui $`\mathrm{\Psi }`$ pentru cazul electronului hidrogenic, trebuie sฤ stabilim unele rechizite generale (care trebuie sฤ se respecte รฎn orice situaลฃie). รn primul rรฎnd, pentru cฤ $`\mathrm{\Psi }^2`$ este proporลฃional cu probabilitatea P de a gฤsi particula descrisฤ prin $`\mathrm{\Psi }`$, integrala $`\mathrm{\Psi }^2`$ pe tot spaลฃiul trebuie sฤ fie finitฤ , pentru ca รฎntr-adevฤr particula sฤ poatฤ fi localizatฤ . Deasemenea, dacฤ
$$_{\mathrm{}}^{\mathrm{}}\mathrm{\Psi }^2๐V=0$$
(1)
particula nu existฤ , iar dacฤ integrala este $`\mathrm{}`$ nu putem avea semnificaลฃie fizicฤ ; $`\mathrm{\Psi }^2`$ nu poate fi negativฤ sau complexฤ din simple motive matematice, astfel cฤ unica posibilitate rฤmรฎne ca integrala sฤ fie finitฤ pentru a avea o descriere acceptabilฤ a unei particule reale. รn general, este convenabil de a identifica $`\mathrm{\Psi }^2`$ cu probabilitatea P de a gฤsi particula descrisฤ de cฤtre $`\mathrm{\Psi }`$ ลi nu doar simpla proporลฃionalitate cu P. Pentru ca $`\mathrm{\Psi }^2`$ sฤ fie egalฤ cu P se impune
$$_{\mathrm{}}^{\mathrm{}}\mathrm{\Psi }^2๐V=1,$$
(2)
pentru cฤ
$$_{\mathrm{}}^{\mathrm{}}\mathrm{P}๐V=1$$
(3)
este afirmaลฃia matematicฤ a faptului cฤ particula existฤ รฎntr-un loc din spaลฃiu la orice moment. O funcลฃie care respectฤ ec. 2 se spune cฤ este normalizatฤ . Pe lรฎngฤ aceastฤ condiลฃie fundamentalฤ , $`\mathrm{\Psi }`$ trebuie sฤ aibฤ o valoare unicฤ , pentru cฤ P are o singurฤ valoare รฎntr-un loc ลi la un moment determinat. O altฤ condiลฃie pe care $`\mathrm{\Psi }`$ trebuie sฤ o satisfacฤ este cฤ atรฎt ea cรฎt ลi derivatele sale parลฃiale $`\frac{\mathrm{\Psi }}{x}`$, $`\frac{\mathrm{\Psi }}{y}`$, $`\frac{\mathrm{\Psi }}{z}`$ trebuie sฤ fie continue รฎn orice punct arbitrar.
Ecuaลฃia Schrรถdinger este consideratฤ ecuaลฃia fundamentalฤ a mecanicii cuantice รฎn acelaลi sens รฎn care legea forลฃei este ecuaลฃia fundamentalฤ a mecanicii newtoniene cu deosebirea importantฤ cฤ este o ecuaลฃie de undฤ pentru $`\mathrm{\Psi }`$.
Odatฤ ce energia potenลฃialฤ este cunoscutฤ , se poate rezolva ecuaลฃia Schrรถdinger pentru funcลฃia de undฤ $`\mathrm{\Psi }`$ a particulei, a cฤrei densitate de probabilitate $`\mathrm{\Psi }^2`$ se poate determina pentru $`x,y,z,t`$. รn multe situaลฃii, energia potenลฃialฤ a unei particule nu depinde explicit de timp; forลฃele care acลฃioneazฤ asupra ei se schimbฤ รฎn funcลฃie numai de posiลฃia particulei. รn aceste condiลฃii, ecuaลฃia Schrรถdinger se poate simplifica eliminรฎnd tot ce se referฤ la $`t`$. Sฤ notฤm cฤ se poate scrie funcลฃia de undฤ unidimensionalฤ a unei particule libere รฎn forma
$`\mathrm{\Psi }(x,t)`$ $`=`$ $`Ae^{(i/\mathrm{})(Etpx)}`$ (4)
$`=`$ $`Ae^{(iE/\mathrm{})t}e^{(ip/\mathrm{})x}`$
$`=`$ $`\psi (x)e^{(iE/\mathrm{})t}.`$
$`\mathrm{\Psi }(x,t)`$ este produsul รฎntre o funcลฃie dependentฤ de timp $`e^{(iE/\mathrm{})t}`$ ลi una staลฃionarฤ , dependentฤ numai de poziลฃie $`\psi (x)`$.
รn cazul general รฎnsฤ , ecuaลฃia Schrรถdinger pentru o stare staลฃionarฤ se poate rezolva numai pentru anumite valori ale energiei E. Nu este vorba de dificultฤลฃi matematice, ci de un aspect fundamental. โA rezolvaโ ecuaลฃia Schrรถdinger pentru un sistem dat รฎnseamnฤ a obลฃine o funcลฃie de undฤ $`\psi `$ care nu numai cฤ satisface ecuaลฃia ลi condiลฃiile de frontierฤ impuse, ci este o funcลฃie de undฤ acceptabilฤ , respectiv, funcลฃia ลi derivata sa sฤ fie continue, finite ลi univoce. Astfel, cuantizarea energiei apare รฎn mecanica ondulatorie ca un element teoretic natural, iar รฎn practicฤ ca un fenomen universal, caracteristic tuturor sistemelor microscopice stabile .
## Ecuaลฃia Schrรถdinger pentru atomul de hidrogen
รn continuare, vom aplica ecuaลฃia Schrรถdinger atomului de hidrogen, despre care se ลtia pe baza experimentelor Rutherford cฤ este format dintr-un proton, particulฤ cu sarcina electricฤ +$`e`$ ลi un electron de sarcinฤ -$`e`$ ลi care fiind de 1836 de ori mai uลor decรฎt protonul este cu mult mai mobil.
Dacฤ interacลฃiunea รฎntre douฤ particule este de tipul $`u(r)=u(\stackrel{}{r}_1\stackrel{}{r}_2)`$, problema de miลcare se reduce atรฎt clasic cรฎt ลi cuantic la miลcarea unei singure particule รฎn cรฎmpul de simetrie sfericฤ . รntr-adevฤr Lagrangeanul:
$$L=\frac{1}{2}m_1\dot{\stackrel{}{r}_1^2}+\frac{1}{2}m_2\dot{\stackrel{}{r}_2^2}u(\stackrel{}{r}_1\stackrel{}{r}_2)$$
(5)
se transformฤ folosind:
$$\stackrel{}{r}=\stackrel{}{r}_1\stackrel{}{r}_2$$
(6)
ลi
$$\stackrel{}{R}=\frac{m_1\stackrel{}{r}_1+m_2\stackrel{}{r}_2}{m_1+m_2},$$
(7)
รฎn Lagrangeanul:
$$L=\frac{1}{2}M\dot{\stackrel{}{R}^2}+\frac{1}{2}\mu \dot{\stackrel{}{r}^2}u(r),$$
(8)
unde
$$M=m_1+m_2$$
(9)
ลi
$$\mu =\frac{m_1m_2}{m_1+m_2}.$$
(10)
Pe de altฤ parte, introducerea impulsului se face cu formulele Lagrange
$$\stackrel{}{P}=\frac{L}{\dot{\stackrel{}{R}}}=M\dot{\stackrel{}{R}}$$
(11)
ลi
$$\stackrel{}{p}=\frac{L}{\dot{\stackrel{}{r}}}=m\dot{\stackrel{}{r}},$$
(12)
ceea ce permite scrierea funcลฃiei Hamilton clasice
$$H=\frac{P^2}{2M}+\frac{p^2}{2m}+u(r).$$
(13)
Astfel, se poate obลฃine operatorul hamiltonian pentru problema corespunzฤtoare cuanticฤ cu comutatori de tipul
$$[P_i,P_k]=i\mathrm{}\delta _{ik}$$
(14)
ลi
$$[p_i,p_k]=i\mathrm{}\delta _{ik}.$$
(15)
Aceลti comutatori implicฤ un operator Hamiltonian de forma
$$\widehat{H}=\frac{\mathrm{}^2}{2M}_R^2\frac{\mathrm{}^2}{2m}_r^2+u(r),$$
(16)
care este fundamental pentru studiul atomului de hidrogen cu ajutorul ecuaลฃiei Schrรถdinger รฎn forma staลฃionarฤ
$$\widehat{H}\psi =E\psi ,$$
(17)
ceea ce presupune cฤ nu se iau รฎn considerare efecte relativiste (viteze apropriate de cele ale luminii รฎn vid).
Energia potenลฃialฤ $`u(r)`$ este cea electrostaticฤ
$$u=\frac{e^2}{4\pi ฯต_0r}$$
(18)
Existฤ douฤ posibilitฤลฃi: prima, de a exprima $`u`$ รฎn funcลฃie de coordonatele carteziene $`x,y,z`$ substituind $`r`$ prin $`\sqrt{x^2+y^2+z^2}`$, a doua, de a exprima ecuaลฃia Schrรถdinger รฎn funcลฃie de coordonatele polare sferice $`r,\theta ,\varphi `$. รn virtutea simetriei sferice a situaลฃiei fizice, vom trata ultimul caz pentru cฤ problema matematicฤ se simplificฤ considerabil.
Prin urmare, รฎn coordonate polare sferice, ecuaลฃia Schrรถdinger este
$$\frac{1}{r^2}\frac{}{r}\left(r^2\frac{\psi }{r}\right)+\frac{1}{r^2\mathrm{sin}\theta }\frac{}{\theta }\left(\mathrm{sin}\theta \frac{\psi }{\theta }\right)+\frac{1}{r^2\mathrm{sin}^2\theta }\frac{^2\psi }{\varphi ^2}+\frac{2m}{\mathrm{}^2}(Eu)\psi =0$$
(19)
Substituind (18) ลi multiplicรฎnd toatฤ ecuaลฃia cu $`r^2\mathrm{sin}^2\theta `$, se obลฃine
$$\mathrm{sin}^2\theta \frac{}{r}\left(r^2\frac{\psi }{r}\right)+\mathrm{sin}\theta \frac{}{\theta }\left(\mathrm{sin}\theta \frac{\psi }{\theta }\right)+\frac{^2\psi }{\varphi ^2}+\frac{2mr^2\mathrm{sin}^2\theta }{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)\psi =0.$$
(20)
Aceastฤ ecuaลฃie este ecuaลฃia diferenลฃialฤ cu derivate parลฃiale pentru funcลฃia de undฤ $`\psi (r,\theta ,\varphi )`$ a electronului รฎn atomul de hidrogen. รmpreunฤ cu diferitele condiลฃii pe care $`\psi (r,\theta ,\varphi )`$ trebuie sฤ le รฎndeplineascฤ \[de exemplu, $`\psi (r,\theta ,\varphi )`$ trebuie sฤ aibฤ o valoare unicฤ pentru fiecare punct spaลฃial ($`r,\theta ,\varphi `$)\], aceastฤ ecuaลฃie specificฤ de manierฤ completฤ comportamentul electronului hidrogenic. Pentru a vedea care este acest comportament, vom rezolva ec. 20 pentru $`\psi (r,\theta ,\varphi )`$ ลi vom interpreta rezultatele obลฃinute.
## Separarea de variabile รฎn coordonate sferice
Ceea ce este cu adevฤrat util รฎn scrierea ecuaลฃiei Schrรถdinger รฎn coordonate sferice pentru atomul de hidrogen constฤ รฎn faptul cฤ astfel se poate realiza uลor separarea รฎn trei ecuaลฃii independente, fiecare unidimensionalฤ . Procedeul de separare este de a cฤuta soluลฃiile pentru care funcลฃia de undฤ $`\psi (r,\theta ,\varphi )`$ are forma unui produs de trei funcลฃii, fiecare รฎntr-una din cele trei variabile sferice: $`R(r)`$, care depinde numai de $`r`$; $`\mathrm{\Theta }(\theta )`$ care depinde numai de $`\theta `$; ลi $`\mathrm{\Phi }(\varphi )`$ care depinde numai de $`\varphi `$ ลi este practic analog separฤrii ecuaลฃiei Laplace. Deci
$$\psi (r,\theta ,\varphi )=R(r)\mathrm{\Theta }(\theta )\mathrm{\Phi }(\varphi ).$$
(21)
Funcลฃia $`R(r)`$ descrie variaลฃia funcลฃiei de undฤ $`\psi `$ a electronului de-a lungul razei vectoare dinspre nucleu, cu $`\theta `$ ลi $`\varphi `$ constante. Variaลฃia lui $`\psi `$ cu unghiul zenital $`\theta `$ de-a lungul unui meridian al unei sfere centratฤ รฎn nucleu este descrisฤ numai de cฤtre funcลฃia $`\mathrm{\Theta }(\theta )`$ pentru $`r`$ ลi $`\varphi `$ constante. รn sfรฎrลit, funcลฃia $`\mathrm{\Phi }(\varphi )`$ descrie cum variazฤ $`\psi `$ cu unghiul azimutal $`\varphi `$ de-a lungul unei paralele a unei sfere centratฤ รฎn nucleu, รฎn condiลฃiile รฎn care $`r`$ ลi $`\theta `$ sunt menลฃinute constante.
Folosind $`\psi =R\mathrm{\Theta }\mathrm{\Phi }`$, vedem cฤ
$$\frac{\psi }{r}=\mathrm{\Theta }\mathrm{\Phi }\frac{dR}{dr},$$
(22)
$$\frac{\psi }{\theta }=R\mathrm{\Phi }\frac{d\mathrm{\Theta }}{d\theta },$$
(23)
$$\frac{\psi }{\varphi }=R\mathrm{\Theta }\frac{d\mathrm{\Phi }}{d\varphi }.$$
(24)
Evident, acelaลi tip de formule se menลฃine pentru derivatele de ordin superior nemixte. Subtituindu-le รฎn ec. 20, dupฤ รฎmpฤrลฃirea cu $`R\mathrm{\Theta }\mathrm{\Phi }`$ se obลฃine
$$\frac{\mathrm{sin}^2\theta }{R}\frac{d}{dr}\left(r^2\frac{dR}{dr}\right)+\frac{\mathrm{sin}\theta }{\mathrm{\Theta }}\frac{d}{d\theta }\left(\mathrm{sin}\theta \frac{d\mathrm{\Theta }}{d\theta }\right)+\frac{1}{\mathrm{\Phi }}\frac{d^2\mathrm{\Phi }}{d\varphi ^2}+\frac{2mr^2\mathrm{sin}^2\theta }{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)=0.$$
(25)
Al treilea termen al acestei ecuaลฃii este funcลฃie numai de unghiul $`\varphi `$, รฎn timp ce ceilalลฃi doi sunt funcลฃii de $`r`$ ลi $`\theta `$. Rescriem ecuaลฃia anterioarฤ รฎn forma
$$\frac{\mathrm{sin}^2\theta }{R}\frac{}{r}\left(r^2\frac{R}{r}\right)+\frac{\mathrm{sin}\theta }{\mathrm{\Theta }}\frac{}{\theta }\left(\mathrm{sin}\theta \frac{\mathrm{\Theta }}{\theta }\right)+\frac{2mr^2\mathrm{sin}^2\theta }{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)=\frac{1}{\mathrm{\Phi }}\frac{^2\mathrm{\Phi }}{\varphi ^2}.$$
(26)
Aceastฤ ecuaลฃie poate fi corectฤ numai dacฤ cei doi membri sunt egali cu aceeaลi constantฤ , pentru cฤ sunt funcลฃii de variabile diferite. Este convenabil sฤ notฤm aceastฤ constantฤ cu $`m_l^2`$. Ecuaลฃia diferenลฃialฤ pentru funcลฃia $`\mathrm{\Phi }`$ este
$$\frac{1}{\mathrm{\Phi }}\frac{^2\mathrm{\Phi }}{\varphi ^2}=m_l^2.$$
(27)
Dacฤ se subtituie $`m_l^2`$ รฎn partea dreaptฤ a ec. 26 ลi se divide ecuaลฃia rezultantฤ cu $`\mathrm{sin}^2\theta `$, dupฤ o regrupare a termenilor, se obลฃine
$$\frac{1}{R}\frac{d}{dr}\left(r^2\frac{dR}{dr}\right)+\frac{2mr^2}{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)=\frac{m_l^2}{\mathrm{sin}^2\theta }\frac{1}{\mathrm{\Theta }\mathrm{sin}\theta }\frac{d}{d\theta }\left(\mathrm{sin}\theta \frac{d\mathrm{\Theta }}{d\theta }\right).$$
(28)
รncฤ odatฤ se prezintฤ o ecuaลฃie รฎn care apar variabile diferite รฎn fiecare membru, ceea ce obligฤ la egalarea ambilor cu aceeaลi constantฤ . Din motive care se vor vedea mai tรฎrziu, vom nota aceastฤ constantฤ prin $`l(l+1)`$. Ecuaลฃiile pentru funcลฃiile $`\mathrm{\Theta }(\theta )`$ ลi $`R(r)`$ sunt
$$\frac{m_l^2}{\mathrm{sin}^2\theta }\frac{1}{\mathrm{\Theta }\mathrm{sin}\theta }\frac{d}{d\theta }\left(sin\theta \frac{d\mathrm{\Theta }}{d\theta }\right)=l(l+1)$$
(29)
ลi
$$\frac{1}{R}\frac{d}{dr}\left(r^2\frac{dR}{dr}\right)+\frac{2mr^2}{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)=l(l+1).$$
(30)
Ecuaลฃiile 27, 29 ลi 30 se scriu รฎn mod normal รฎn forma
$$\frac{d^2\mathrm{\Phi }}{d\varphi ^2}+m_l^2\mathrm{\Phi }=0,$$
(31)
$$\frac{1}{\mathrm{sin}\theta }\frac{d}{d\theta }\left(\mathrm{sin}\theta \frac{d\mathrm{\Theta }}{d\theta }\right)+\left[l(l+1)\frac{m_l^2}{\mathrm{sin}^2\theta }\right]\mathrm{\Theta }=0,$$
(32)
$$\frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{dR}{dr}\right)+\left[\frac{2m}{\mathrm{}^2}\left(\frac{e^2}{4\pi ฯต_0r}+E\right)\frac{l(l+1)}{r^2}\right]R=0.$$
(33)
Fiecare dintre aceste ecuaลฃii este o ecuaลฃie diferenลฃialฤ ordinarฤ pentru o funcลฃie de o singurฤ variabilฤ . รn felul acesta s-a reuลit simplificarea ecuaลฃiei Schrรถdinger pentru atomul de hidrogen care, iniลฃial, era o ecuaลฃie diferenลฃialฤ parลฃialฤ pentru o funcลฃie $`\psi `$ de trei variabile.
## Interpretarea constantelor de separare: numere cuantice
### Soluลฃia pentru partea azimutalฤ
Ec. 31 se rezolvฤ uลor pentru a gฤsi urmฤtoarea soluลฃie
$$\mathrm{\Phi }(\varphi )=A_\varphi e^{im_l\varphi },$$
(34)
unde $`A_\varphi `$ este constanta de integrare. Una dintre condiลฃiile stabilite mai รฎnainte pe care trebuie sฤ le รฎndeplineascฤ o funcลฃie de undฤ (ลi prin urmare deasemenea $`\mathrm{\Phi }`$, care este o componentฤ a funcลฃiei complete $`\psi `$) este sฤ aibฤ o valoare unicฤ pentru fiecare punct din spaลฃiu fฤrฤ excepลฃie. De exemplu, se observฤ cฤ $`\varphi `$ ลi $`\varphi +2\pi `$ se identificฤ รฎn acelaลi plan meridian. De aceea, trebuie ca $`\mathrm{\Phi }(\varphi )=\mathrm{\Phi }(\varphi +2\pi )`$, adicฤ $`Ae^{im_l\varphi }=Ae^{im_l(\varphi +2\pi )}`$. Aceasta se poate รฎndeplini numai cรฎnd $`m_l`$ este 0 sau un numฤr รฎntreg pozitiv sau negativ $`(\pm 1,\pm 2,\pm 3,\mathrm{})`$. Acest numฤr $`m_l`$ se cunoaลte ca numฤrul cuantic magnetic al atomului de hidrogen ลi este relaลฃionat cu direcลฃia momentului cinetic $`L`$ pentru cฤ s-a putut fi asociat cu efectele cรฎmpurilor magnetice axiale asupra electronului. Numฤrul cuantic magnetic $`m_l`$ este determinat de cฤtre numฤrul cuantic orbital $`l`$, care la rรฎndul sฤu determinฤ modulul momentului cinetic al electronului.
Interpretarea numฤrului cuantic orbital $`l`$ nu este nici ea fฤrฤ unele probleme. Sฤ examinฤm ec. 33, care corespunde pฤrลฃii radiale $`R(r)`$ a funcลฃiei de undฤ $`\psi `$. Aceastฤ ecuaลฃie este relaลฃionatฤ numai cu aspectul radial al miลcฤrii electronilor, adicฤ , cu apropierea ลi depฤrtarea de nucleu (pentru elipse); totuลi, este prezentฤ ลi energia totalฤ a electronului $`E`$. Aceastฤ energie include energia cineticฤ a electronului รฎn miลcare orbitalฤ care nu are nimic de-a face cu miลcarea radialฤ . Aceastฤ contradicลฃie se poate elimina cu urmฤtorul raลฃionament: energia cineticฤ $`T`$ a electronului are douฤ pฤrลฃi: $`T_{radial}`$ datoratฤ miลcฤrii de apropiere ลi depฤrtare de nucleu, ลi $`T_{orbital}`$ datoratฤ miลcฤrii รฎn jurul nucleului. Energia potenลฃialฤ $`V`$ a electronului este energia electrostaticฤ . Prin urmare, energia sa totalฤ este
$$E=T_{radial}+T_{orbital}\frac{e^2}{4\pi ฯต_0r}.$$
(35)
Substituind aceastฤ expresie a lui $`E`$ รฎn ec. 33 obลฃinem, dupฤ o regrupare a termenilor,
$$\frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{dR}{dr}\right)+\frac{2m}{\mathrm{}^2}\left[T_{radial}+T_{orbital}\frac{\mathrm{}^2l(l+1)}{2mr^2}\right]R=0.$$
(36)
Dacฤ ultimii doi termeni din paranteze se anuleazฤ รฎntre ei, obลฃinem o ecuaลฃie diferenลฃialฤ pentru miลcarea pur radialฤ . Impunem deci condiลฃia
$$T_{orbital}=\frac{\mathrm{}^2l(l+1)}{2mr^2}.$$
(37)
Energia cineticฤ orbitalฤ a electronului este รฎnsฤ
$$T_{orbital}=\frac{1}{2}mv_{orbital}^2$$
(38)
ลi cum momentul cinetic $`L`$ al electronului este
$$L=mv_{orbital}r,$$
(39)
putem exprima energia cineticฤ orbitalฤ รฎn forma
$$T_{orbital}=\frac{L^2}{2mr^2}$$
(40)
De aceea avem
$$\frac{L^2}{2mr^2}=\frac{\mathrm{}^2l(l+1)}{2mr^2}$$
(41)
ลi deci
$$L=\sqrt{l(l+1)}\mathrm{}.$$
(42)
Interpretarea acestui rezultat este cฤ , รฎntrucรฎt numฤrul cuantic orbital $`l`$ este limitat la valorile $`l=0,1,2,\mathrm{},(n1)`$, electronul poate avea numai momentele cinetice $`L`$ care se specificฤ prin intermediul ec. 42. Ca ลi รฎn cazul energiei totale $`E`$, momentul cinetic se conservฤ ลi este cuantizat, iar unitatea sa naturalฤ de mฤsurฤ รฎn mecanica cuanticฤ este $`\mathrm{}=h/2\pi =1.054\times 10^{34}`$ J.s.
รn miลcarea planetarฤ macroscopicฤ , numฤrul cuantic care descrie momentul unghiular este atรฎt de mare cฤ separarea รฎn stฤri discrete ale momentului cinetic nu se poate observa experimental. De exemplu, un electron al cฤrui numฤr cuantic orbital este 2, are un moment cinetic $`L=2.6\times 10^{34}`$ J.s., รฎn timp ce momentul cinetic al planetei noastre este $`2.7\times 10^{40}`$ J.s.!
Se obiลnuieลte sฤ se noteze stฤrile de moment cinetic cu litera $`s`$ pentru $`l=0`$, cu $`p`$ pentru $`l=1`$, $`d`$ pentru $`l=2`$, etc. Acest cod alfabetic provine din clasificarea empiricฤ a spectrelor รฎn aลa numitele serii principalฤ , difuzฤ ลi fundamentalฤ , care este anterioarฤ mecanicii cuantice.
Combinarea numฤrului cuantic total cu litera corespunzฤtoare momentului cinetic este o altฤ notaลฃie frecvent folositฤ pentru stฤrile atomice. De exemplu, o stare รฎn care $`n=2`$, $`l=0`$ este o stare $`2s`$, iar una รฎn care $`n=4`$, $`l=2`$ este o stare $`4d`$.
Pe de altฤ parte, pentru interpretarea numฤrului cuantic magnetic, vom ลฃine cont cฤ la fel ca pentru impulsul lineal, momentul cinetic este un vector ลi deci pentru al descrie se necesitฤ specificarea direcลฃiei, sensului ลi modulului sฤu. Vectorul $`L`$ este perpendicular planului รฎn care are loc miลcarea de rotaลฃie ลi direcลฃia ลi sensul sฤu sunt date de regula mรฎinii drepte (de produs vectorial): degetul mare are direcลฃia ลi sensul lui $`L`$ cรฎnd celelalte patru degete sunt รฎn direcลฃia de rotaลฃie.
Dar ce semnificaลฃie se poate da unei direcลฃii ลi sens รฎn spaลฃiul limitat al unui atom de hidrogen ? Rฤspunsul este simplu dacฤ ne gรฎndim cฤ un electron care gireazฤ รฎn jurul unui nucleu reprezintฤ un circuit minuscul, care ca dipol magnetic prezintฤ un cรฎmp magnetic corespunzฤtor. รn consecinลฃฤ , un electron atomic cu moment cinetic interacลฃioneazฤ cu un cรฎmp magnetic extern $`B`$. Numฤrul cuantic magnetic $`m_l`$ specificฤ direcลฃia lui $`L`$, determinatฤ de componenta lui $`L`$ รฎn direcลฃia cรฎmpului. Acest fenomen se cunoaลte รฎn mod comun drept cuantizare spaลฃialฤ .
Dacฤ alegem direcลฃia cรฎmpului magnetic ca axฤ $`z`$, componenta lui $`L`$ รฎn aceastฤ direcลฃie este
$$L_z=m_l\mathrm{}.$$
(43)
Valorile posibile ale lui $`m_l`$ pentru o valoare datฤ a lui $`l`$, merg de la $`+l`$ pรฎnฤ la $`l`$, trecรฎnd prin 0, astfel cฤ orientฤrile posibile ale vectorului moment cinetic $`L`$ รฎntr-un cรฎmp magnetic sunt $`2l+1`$. Cรฎnd $`l=0`$, $`L_z`$ poate avea numai valoarea zero; cรฎnd $`l=1`$, $`L_z`$ poate fi $`\mathrm{}`$, 0, sau $`\mathrm{}`$; cรฎnd $`l=2`$, $`L_z`$ ia numai una dintre valorile $`2\mathrm{}`$, $`\mathrm{}`$, 0, $`\mathrm{}`$, sau $`2\mathrm{}`$, ลi aลa mai departe. Menลฃionฤm cฤ $`L`$ nu poate fi exact alineat (paralel sau antiparalel) cu $`B`$, pentru cฤ $`L_z`$ este รฎntotdeauna mai mic decรฎt modulul $`\sqrt{l(l+1)}\mathrm{}`$ momentului unghiular total.
Cuantizarea spaลฃialฤ a momentului cinetic orbital al atomului de hidrogen se aratฤ รฎn fig. 6.1.
Fig. 6.1: Cuantizarea spaลฃialฤ a momentului cinetic pentru stฤri $`l=2`$, $`L=\sqrt{6}\mathrm{}`$.
Trebuie sฤ considerฤm electronul caracterizat de cฤtre un anumit $`m_l`$ ca avรฎnd o orientare determinatฤ a momentului sฤu cinetic $`L`$ faลฃฤ de un cรฎmp magnetic extern รฎn cazul รฎn care acesta se aplicฤ .
รn absenลฃa cรฎmpului magnetic extern, direcลฃia axei $`z`$ este complet arbitrarฤ . De aceea, componenta lui $`L`$ รฎn orice direcลฃie pe care o alegem este $`m_l\mathrm{}`$; cรฎmpul magnetic extern oferฤ o direcลฃie de referinลฃฤ privilegiatฤ din punct de vedere experimental.
Dece este cuantizatฤ numai componenta lui $`L`$ ? Rฤspunsul se relaลฃioneazฤ cu faptul cฤ $`L`$ nu poate fi direcลฃionat de manierฤ arbitrarฤ ; รฎntotdeauna descrie un con centrat pe axa de cuantizare รฎn aลa fel รฎncรฎt proiecลฃia sa $`L_z`$ este $`m_l\mathrm{}`$. Motivul pentru care se produce acest fenomen este principiul de incertitudine: dacฤ $`L`$ ar fi fix รฎn spaลฃiu, รฎn aลa fel รฎncรฎt $`L_x`$, $`L_y`$ ลi $`L_z`$ ar avea valori bine definite, electronul ar fi confinat รฎntr-un plan bine definit. De exemplu, dacฤ $`L`$ ar fi fixat de-a lungul direcลฃiei $`z`$, electronul ar avea tendinลฃa de a se menลฃine รฎn planul $`xy`$ (fig. 6.2a).
Fig. 6.2: Principiul de incertitudine interzice o direcลฃie fixฤ รฎn spaลฃiu a momentului cinetic.
Acest lucru poate sฤ aibฤ loc numai รฎn situaลฃia รฎn care componenta $`p_z`$ a impulsului electronului รฎn direcลฃia $`z`$ este infinit de incertฤ , ceea ce desigur este imposibil dacฤ face parte din atomul de hidrogen. Totuลi, cum รฎn realitate numai o componentฤ $`L_z`$ a lui $`L`$ รฎmpreunฤ cu $`L^2`$ au valori definite ลi $`L>L_z`$, electronul nu este limitat la un plan unic (fig. 6.2b), iar dacฤ ar fi aลa, ar exista o incertitudine รฎn coordonata $`z`$ a electronului. Direcลฃia lui $`L`$ se schimbฤ รฎn mod constant (fig. 6.3), astfel cฤ valorile medii ale lui $`L_x`$ ลi $`L_y`$ sunt 0, deลi $`L_z`$ are รฎntotdeauna valoarea $`m_l\mathrm{}`$.
Fig. 6.3: Vectorul moment cinetic prezintฤ o precesie constantฤ รฎn jurul axei $`z`$.
Soluลฃia pentru $`\mathrm{\Phi }`$ trebuie sฤ satisfacฤ deasemenea condiลฃia de normalizare, care este datฤ de cฤtre ec. 2. Deci pentru $`\mathrm{\Phi }`$ avem
$$_0^{2\pi }\mathrm{\Phi }^2๐\varphi =1$$
(44)
ลi substituind $`\mathrm{\Phi }`$ se obลฃine
$$_0^{2\pi }A_\varphi ^2๐\varphi =1.$$
(45)
Astfel $`A_\varphi =1/\sqrt{2\pi }`$ ลi deci $`\mathrm{\Phi }`$ normalizatฤ este datฤ de
$$\mathrm{\Phi }(\varphi )=\frac{1}{\sqrt{2\pi }}e^{im_l\varphi }.$$
(46)
### Soluลฃia pentru partea polarฤ
Ecuaลฃia diferenลฃialฤ pentru partea polarฤ $`\mathrm{\Theta }(\theta )`$ are o soluลฃie mai complicatฤ fiind datฤ de polinoamele Legendre asociate
$$P_l^{m_l}(x)=(1)^{m_l}(1x^2)^{m_l/2}\frac{d^{m_l}}{dx^{m_l}}P_l(x)=(1)^{m_l}\frac{(1x^2)^{m_l/2}}{2^ll!}\frac{d^{m_l+l}}{dx^{m_l+l}}(x^21)^l.$$
(47)
Aceste funcลฃii satisfac urmฤtoarea relaลฃie de ortogonalitate
$$_1^1[P_l^{m_l}(cos\theta )]^2๐cos\theta =\frac{2}{2l+1}\frac{(l+m_l)!}{(lm_l)!}.$$
(48)
รn cazul mecanicii cuantice, $`\mathrm{\Theta }(\theta )`$ este datฤ de polinoamele Legendre normalizate, respectiv, dacฤ
$$\mathrm{\Theta }(\theta )=A_\theta P_l^{m_l}(cos\theta ),$$
(49)
atunci condiลฃia de normalizare este datฤ de
$$_1^1A_\theta ^2[P_l^{m_l}(cos\theta )]^2๐cos\theta =1.$$
(50)
Prin urmare constanta de normalizare pentru partea polarฤ este
$$A_\theta =\sqrt{\frac{2l+1}{2}\frac{(lm_l)!}{(l+m_l)!}}$$
(51)
ลi prin urmare, funcลฃia $`\mathrm{\Theta }(\theta )`$ deja normalizatฤ este
$$\mathrm{\Theta }(\theta )=\sqrt{\frac{2l+1}{2}\frac{(lm_l)!}{(l+m_l)!}}P_l^{m_l}(cos\theta )$$
(52)
Pentru obiectivele noastre, cea mai importantฤ proprietate a acestor funcลฃii este cฤ , aลa cum s-a menลฃionat deja, existฤ numai cรฎnd constanta $`l`$ este un numฤr รฎntreg egal sau mai mare decรฎt $`m_l`$, care este valoarea absolutฤ a lui $`m_l`$. Aceastฤ condiลฃie se poate scrie sub forma setului de valori disponibile pentru $`m_l`$
$$m_l=0,\pm 1,\pm 2,\mathrm{},\pm l.$$
(53)
### Unificarea pฤrลฃilor azimutalฤ ลi polarฤ : armonicele sferice
Soluลฃiile pentru pฤrลฃile azimutalฤ ลi polarฤ se pot uni pentru a forma armonicele sferice, care depind de $`\varphi `$ ลi $`\theta `$ ลi contribuie la simplificarea manipulฤrilor algebrice ale funcลฃiei de undฤ completฤ $`\psi (r,\theta ,\varphi )`$. Armonicele sferice se introduc รฎn felul urmฤtor:
$$Y_l^{m_l}(\theta ,\varphi )=(1)^{m_l}\sqrt{\frac{2l+1}{4\pi }\frac{(lm_l)!}{(l+m_l)!}}P_l^{m_l}(cos\theta )e^{im_l\varphi }.$$
(54)
Factorul suplimentar $`(1)^{m_l}`$ nu produce nici o problemฤ pentru cฤ ec. Schrรถdinger este linearฤ ลi homogenฤ ลi este convenabil pentru studiul del momentului cinetic. Se cunoaลte ca factorul de fazฤ Condon-Shortley, efectul sฤu fiind de a introduce o alternanลฃฤ a semnelor $`\pm `$.
### Soluลฃia pentru partea radialฤ
Soluลฃia pentru partea radialฤ $`R(r)`$ a funcลฃiei de undฤ $`\psi `$ a atomului de hidrogen este ceva mai complicatฤ ลi aici este unde apar diferenลฃe mai mari faลฃฤ de ecuaลฃia Laplace รฎn electrostaticฤ . Rezultatul final se exprimฤ รฎn funcลฃie de polinoamele asociate Laguerre (Schrรถdinger 1926). Ecuaลฃia radialฤ se poate rezolva รฎn formฤ analiticฤ exactฤ numai cรฎnd E este pozitiv sau pentru una din urmฤtoarele valori negative $`E_n`$ (รฎn care caz electronul este legat atomului)
$$E_n=\frac{me^4}{32\pi ^2ฯต_0^2\mathrm{}^2}\left(\frac{1}{n^2}\right),$$
(55)
unde $`n`$ este un numฤr รฎntreg numit numฤrul cuantic principal ลi descrie cuantizarea energiei electronului รฎn atomul de hidrogen. Aceast spectru discret a fost obลฃinut pentru prima datฤ de cฤtre Bohr cu metode empirice de cuantizare รฎn 1913 ลi apoi de cฤtre Pauli ลi respectiv Schrรถdinger รฎn 1926.
O altฤ condiลฃie care trebuie sฤ fie satisfฤcutฤ pentru a rezolva ecuaลฃia radialฤ este ca $`n`$ sฤ fie รฎntotdeauna mai mare decรฎt $`l`$. Valoarea sa minimฤ este $`l+1`$. Invers, condiลฃia asupra lui $`l`$ este
$$l=0,1,2,\mathrm{},(n1)$$
(56)
Ecuaลฃia radialฤ se poate scrie รฎn forma
$$r^2\frac{d^2R}{dr^2}+2r\frac{dR}{dr}+\left[\frac{2mE}{\mathrm{}^2}r^2+\frac{2me^2}{4\pi ฯต_0\mathrm{}^2}rl(l+1)\right]R=0,$$
(57)
Dupฤ รฎmpฤrลฃirea cu $`r^2`$, se foloseลte substituลฃia $`\chi (r)=rR`$ pentru a elimina termenul รฎn $`\frac{dR}{dr}`$ ลi a obลฃine forma standard a ec. Schrรถdinger radiale cu potenลฃial efectiv $`U(r)=\mathrm{const}/r+l(l+1)/r^2`$ (potenลฃial electrostatic plus barierฤ centrifugalฤ ). Aceastฤ procedurฤ se aplicฤ numai pentru a discuta o nouฤ condiลฃie obligatorie de frontierฤ , obลฃinerea spectrului fiind prin intermediul ecuaลฃiei pentru $`R`$. Diferenลฃa รฎntre o ec. Schrรถdinger radialฤ ลi una รฎn toatฤ linia realฤ este cฤ o condiลฃie de frontierฤ suplimentarฤ trebuie impusฤ รฎn origine ($`r=0`$). Potenลฃialul coulombian aparลฃine unei clase de potenลฃiale care se numesc slab singulare, pentru care $`\mathrm{lim}_{r0}=U(r)r^2=0`$. Se รฎncearcฤ soluลฃii de tipul $`\chi r^\nu `$, ceea ce implicฤ $`\nu (\nu 1)=l(l+1)`$ cu soluลฃiile $`\nu _1=l+1`$ ลi $`\nu _2=l`$, exact ca รฎn cazul electrostaticii. Soluลฃia negativฤ se eliminฤ รฎn cazul $`l0`$ pentru cฤ duce la divergenลฃa integralei de normalizare ลi deasemenea nu respectฤ normalizarea la funcลฃia delta รฎn cazul spectrului continuu, iar cazul $`\nu _2=0`$ se eliminฤ din condiลฃia de finitudine a energiei cinetice medii.Concluzia finalฤ este cฤ $`\chi (0)=0`$ pentru orice $`l`$.
Revenind la analiza ecuaลฃiei pentru funcลฃia radialฤ $`R`$, se pune mai รฎntรฎi problema adimensionalizฤrii ecuaลฃiei. Aceasta se face observรฎnd cฤ se poate forma o singurฤ scalฤ de spaลฃiu ลi timp din combinaลฃii ale celor trei constante fizice care intrฤ รฎn aceastฤ problemฤ respectiv $`e^2`$, $`m`$ ลi $`\mathrm{}`$. Acestea sunt raza Bohr $`a_0=\mathrm{}^2/me^2=0.52910^8`$ cm. ลi $`t_0=\mathrm{}^3/me^4=0.24210^{16}`$ sec., care se numesc unitฤลฃi atomice. Folosind aceste unitฤลฃi obลฃinem
$$\frac{d^2R}{dr^2}+\frac{2}{r}\frac{dR}{dr}+\left[2E+\frac{2}{r}\frac{l(l+1)}{r^2}\right]R=0,$$
(58)
unde ne intereseazฤ spectrul discret ($`E<0`$). Cu notaลฃiile $`n=1/\sqrt{E}`$ ลi $`\rho =2r/n`$ se ajunge la:
$$\frac{d^2R}{d\rho ^2}+\frac{2}{\rho }\frac{dR}{d\rho }+\left[\frac{n}{\rho }\frac{1}{4}\frac{l(l+1)}{\rho ^2}\right]R=0.$$
(59)
Pentru $`\rho \mathrm{}`$, ecuaลฃia se reduce la $`\frac{d^2R}{d\rho ^2}=\frac{R}{4}`$ cu soluลฃii $`Re^{\pm \rho /2}`$. Se acceptฤ pe baza condiลฃiei de normalizare numai exponenลฃiala atenuatฤ . Pe de altฤ parte asimptotica de zero, aลa cum am comentat deja, este $`R\rho ^l`$. Prin urmare, putem substitui $`R`$ printr-un produs de trei funcลฃii radiale $`R=\rho ^le^{\rho /2}F(\rho )`$, dintre care primele douฤ sunt pฤrลฃile asimptotice, iar a treia este funcลฃia radialฤ รฎn regiunea intermediarฤ , care ne intereseazฤ cel mai mult pentru cฤ ne dฤ spectrul energetic. Ecuaลฃia pentru $`F`$ este
$$\rho \frac{d^2F}{d\rho ^2}+(2l+2\rho )\frac{dF}{d\rho }+(nl1)F=0.$$
(60)
care este un caz particular de ecuaลฃie hipergeometricฤ confluentฤ รฎn care cei doi parametri โhiperโgeometrici depind de $`n,l`$ ลi care se poate identifica cu ecuaลฃia pentru polinoamele Laguerre asociate $`L_{n+l}^{2l+1}(\rho )`$ รฎn fizica matematicฤ . Astfel, forma normalizatฤ a lui $`R`$ este:
$$R_{nl}(r)=\frac{2}{n^2}\sqrt{\frac{(nl1)!}{2n[(n+l)!]^3}}e^{\rho /2}\rho ^lL_{n+l}^{2l+1}(\rho ),$$
(61)
unde s-a folosit condiลฃia de normalizare a polinoamelor Laguerre:
$$_0^{\mathrm{}}e^\rho \rho ^{2l}[L_{n+l}^{2l+1}(\rho )]^2\rho ^2๐\rho =\frac{2n[(n+l)!]^3}{(nl1)!}.$$
(62)
Avem deci soluลฃiile fiecฤreia dintre ecuaลฃiile care depind numai de o singurฤ variabilฤ ลi prin urmare putem construi funcลฃia de undฤ pentru fiecare stare electronicฤ รฎn atomul de hidrogen, respectiv dacฤ $`\psi (r,\theta ,\varphi )=R(r)\mathrm{\Theta }(\theta )\mathrm{\Phi }(\varphi )`$, atunci funcลฃia de undฤ completฤ este
$$\psi (r,\theta ,\varphi )=๐ฉ_H(\alpha r)^le^{\alpha r/2}L_{n+l}^{2l+1}(\alpha r)P_l^{m_l}(cos\theta )e^{im_l\varphi },$$
(63)
unde $`๐ฉ_H=\frac{2}{n^2}\sqrt{\frac{2l+1}{4\pi }\frac{(lm_l)!}{(l+m_l)!}\frac{(nl1)!}{[(n+l)!]^3}}`$ ลi $`\alpha =2/na_0`$.
Utilizรฎnd armonicele sferice, soluลฃia se scrie รฎn felul urmฤtor
$$\psi (r,\theta ,\varphi )=\frac{2}{n^2}\sqrt{\frac{(nl1)!}{[(n+l)!]^3}}(\alpha r)^le^{\alpha r/2}L_{n+l}^{2l+1}(\alpha r)Y_l^{m_l}(\theta ,\varphi ).$$
(64)
Aceastฤ formulฤ se poate considera rezultatul matematic final pentru soluลฃia ec. Schrรถdinger รฎn cazul atomului de hidrogen pentru oricare stare staลฃionarฤ a electronului sฤu. รntr-adevฤr, se pot vedea รฎn mod explicit atรฎt dependenลฃa asimptoticฤ cรฎt ลi cele douฤ seturi ortogonale complete, polinoamele Laguerre asociate ลi respectiv armonicele sferice, corespunzฤtoare acestei ecuaลฃii lineare cu derivate parลฃiale de ordinul doi. Coordonatele parabolice \[$`\xi =r(1\mathrm{cos}\theta )`$, $`\eta =r(1+\mathrm{cos}\theta )`$, $`\varphi =\varphi `$\], sunt un alt set de variabile รฎn care ec. Schrรถdinger pentru atomul de hidrogen este uลor de separat (E. Schrรถdinger, Ann. Physik 80, 437, 1926; P.S. Epstein, Phys. Rev. 28, 695, 1926; I. Waller, Zf. Physik 38, 635, 1926). Soluลฃia finalฤ se exprimฤ ca produsul unor factori de naturฤ asimptoticฤ, armonice azimutale ลi douฤ seturi de polinoame Laguerre asociate รฎn $`\xi `$, respectiv $`\eta `$. Spectrul energetic ($`1/n^2`$) ลi degenerarea ($`n^2`$) evident nu se modificฤ.
## Densitatea de probabilitate electronicฤ
รn modelul lui Bohr al atomului de hidrogen, electronul se roteลte รฎn jurul nucleului pe traiectorii circulare sau eliptice. Dacฤ se realizeazฤ un experiment adecuat, s-ar putea vedea cฤ electronul ar fi รฎntotdeauna situat รฎn limitele expeimentale la o distanลฃฤ faลฃฤ de nucleu $`r=n^2a_0`$ (unde $`n`$ este numฤrul cuantic care numeroteazฤ orbita ลi $`a_0=0.53`$ $`\AA `$ este raza orbitei celei mai apropiate de nucleu, cunoscutฤ ca raza Bohr) ลi รฎn planul ecuatorial $`\theta =90^o`$, รฎn timp ce unghiul azimutal $`\varphi `$ poate varia รฎn timp.
Teoria cuanticฤ a atomului de hidrogen modificฤ concluziile modelului lui Bohr รฎn douฤ aspecte importante. รn primul rรฎnd, nu se pot da valori exacte pentru $`r,\theta ,\varphi `$, ci numai probabilitฤลฃi relative de a gฤsi electronul รฎntr-o zonฤ infinitezimalฤ datฤ a spaลฃiului. Aceastฤ imprecizie este, desigur, o consecinลฃฤ a naturii ondulatorii a electronului. รn al doilea rรฎnd, nu se poate gรฎndi cฤ electronul se miลcฤ รฎn jurul nucleului รฎn sensul convenลฃional clasic, pentru cฤ densitatea de probabilitate $`\psi ^2`$ nu depinde de timp ลi poate varia considerabil รฎn funcลฃie de zona infinitezimalฤ unde se calculeazฤ .
Funcลฃia de undฤ $`\psi `$ a electronului รฎn atomul de hidrogen este $`\psi =R\mathrm{\Theta }\mathrm{\Phi }`$ unde $`R=R_{nl}(r)`$ descrie cum se schimbฤ $`\psi `$ cu $`r`$ cรฎnd numerele cuantice orbital ลi total au valorile $`n`$ ลi $`l`$; $`\mathrm{\Theta }=\mathrm{\Theta }_{lm_l}(\theta )`$ descrie la rรฎndul lui variaลฃia lui $`\psi `$ cu $`\theta `$ cรฎnd numerele cuantice magnetic ลi orbital au valorile $`l`$ ลi $`m_l`$; รฎn sfรฎrลit, $`\mathrm{\Phi }=\mathrm{\Phi }_{m_l}(\varphi )`$ dฤ schimbarea lui $`\psi `$ cu $`\varphi `$ cรฎnd numฤrul cuantic magnetic are valoarea $`m_l`$. Densitatea de probabilitate $`\psi ^2`$ se poate scrie
$$\psi ^2=R^2\mathrm{\Theta }^2\mathrm{\Phi }^2.$$
(65)
Densitatea de probabilitate $`\mathrm{\Phi }^2`$, care mฤsoarฤ posibilitatea de a gฤsi electronul la un unghi azimutal $`\varphi `$ dat, este o constantฤ care nu depinde de $`\varphi `$. Prin urmare, densitatea de probabilitate electronicฤ este simetricฤ faลฃฤ de axa $`z`$, independent de starea cuanticฤ โmagneticฤ โ (atรฎta timp cรฎt nu se aplicฤ un cรฎmp magnetic extern), ceea ce face ca electronul sฤ aibฤ aceeaลi probabilitate de a se gฤsi รฎn orice direcลฃie azimutalฤ . Partea radialฤ $`R`$ a funcลฃiei de undฤ , spre deosebire de $`\mathrm{\Phi }`$, nu numai cฤ variazฤ cu $`r`$, ci ลi o face รฎn mod diferit pentru fiecare combinaลฃie de numere cuantice $`n`$ ลi $`l`$. Fig. 6.4 aratฤ grafice ale lui $`R`$ รฎn funcลฃie de $`r`$ pentru stฤrile $`1s`$, $`2s`$, ลi $`2p`$ ale atomului de hidrogen. $`R`$ este maxim รฎn centrul nucleului ($`r=0`$) pentru toate stฤrile $`s`$, รฎn timp ce este zero รฎn $`r=0`$ pentru toate stฤrile care au moment cinetic.
Fig. 6.4: Grafice aproximative ale funcลฃiilor radiale $`R_{1s}`$, $`R_{2s}`$, $`R_{2p}`$; ($`a_0=0.53`$ ร
).
Fig. 6.5: Densitatea de probabilitate de a gฤsi electronul atomului de hidrogen รฎntre $`r`$ ลi $`r+dr`$ faลฃฤ de nucleu pentru stฤrile $`1s`$, $`2s`$, $`2p`$.
Densitatea de probabilitate electronicฤ รฎn punctul $`r,\theta ,\varphi `$ este proporลฃionalฤ cu $`\psi ^2`$, dar probabilitatea realฤ รฎn elementul de volum infinitezimal $`dV`$ este $`\psi ^2dV`$. รn coordonate polare sferice
$$dV=r^2\mathrm{sin}\theta drd\theta d\varphi ,$$
(66)
ลi cum $`\mathrm{\Theta }`$ ลi $`\mathrm{\Phi }`$ sunt funcลฃii normalizate, probabilitatea numericฤ realฤ $`P(r)dr`$ de a gฤsi electronul la o distanลฃฤ faลฃฤ de nucleu cuprinsฤ รฎntre $`r`$ ลi $`r+dr`$ este
$`P(r)dr`$ $`=`$ $`r^2R^2dr{\displaystyle _0^\pi }\mathrm{\Theta }^2\mathrm{sin}\theta d\theta {\displaystyle _0^{2\pi }}\mathrm{\Phi }^2๐\varphi `$ (67)
$`=`$ $`r^2R^2dr`$
$`P(r)`$ este reprezentatฤ รฎn fig. 6.5 pentru aceleaลi stฤri ale cฤror funcลฃii radiale $`R`$ apar รฎn fig. 6.4; รฎn principiu, curbele sunt foarte diferite. Observฤm imediat cฤ $`P(r)`$ nu este maximฤ รฎn nucleu pentru stฤrile $`s`$, aลa cum este $`R`$, avรฎnd maximul la o distanลฃฤ finitฤ de acesta. Valoarea cea mai probabilฤ a lui $`r`$ pentru un electron $`1s`$ este exact $`a_0`$, care este raza Bohr. Totuลi, valoarea medie a lui $`r`$ pentru un electron $`1s`$ este $`1.5a_0`$, ceea ce pare ciudat la prima vedere, pentru cฤ nivelele de energie sunt aceleลi รฎn mecanica cuanticฤ ลi รฎn modelul lui Bohr. Aceastฤ aparentฤ discrepanลฃฤ se eliminฤ dacฤ se ลฃine cont de faptul cฤ energia electronului depinde de $`1/r`$ ลi nu direct de $`r`$, iar valoarea medie a lui $`1/r`$ pentru un electron $`1s`$ este exact $`1/a_0`$.
Funcลฃia $`\mathrm{\Theta }`$ variazฤ cu unghiul polar $`\theta `$ pentru toate numerele cuantice $`l`$ ลi $`m_l`$, excepลฃie fฤcรฎnd $`l=m_l=0`$, care sunt stฤri $`s`$. Densitatea de probabilitate $`\mathrm{\Theta }^2`$ pentru o stare $`s`$ este o constantฤ (1/2), ceea ce รฎnseamnฤ cฤ , รฎntrucรฎt $`\mathrm{\Phi }^2`$ este deasemenea constantฤ , densitatea de probabilitate electronicฤ $`\psi ^2`$ are aceeaลi valoare pentru o valoare a lui $`r`$ datฤ , รฎn toate direcลฃiile. รn alte stฤri, electronii au un comportament unghiular care uneori ajunge sฤ fie foarte complicat. Aceasta se poate vedea รฎn fig.6.5, unde se aratฤ densitฤลฃile de probabilitate electronicฤ pentru diferite stฤri atomice รฎn funcลฃie de $`r`$ ลi $`\theta `$. (Termenul care se reprezintฤ este $`\psi ^2`$ ลi nu $`\psi ^2dV`$). Deoarece $`\psi ^2`$ este independent de $`\varphi `$, o reprezentare tridimensionalฤ a lui $`\psi ^2`$ se obลฃine prin rotaลฃia unei reprezentฤri particulare รฎn jurul unei axe verticale, ceea ce poate arฤta cฤ densitฤลฃile de probabilitate pentru stฤrile $`s`$ au simetrie sfericฤ , รฎn timp ce toate celelalte nu o posedฤ . Se obลฃin รฎn acest fel loburi mai mult sau mai puลฃin pronunลฃate, care au forme caracteristice pentru fiecare stare รฎn parte ลi care รฎn chimie joacฤ un rol important รฎn determinarea modului รฎn care interacลฃioneazฤ atomii รฎn interiorul moleculelor.
6N. Notฤ:
1. E. Schrรถdinger a obลฃinut premiul Nobel รฎn 1933 (รฎmpreunฤ cu Dirac) pentru โdescoperirea de noi forme productive ale teoriei atomiceโ. Schrรถdinger a scris o remarcabilฤ serie de patru articole intitulatฤ โQuantisierung als Eigenwertproblemโ \[โCuantizarea ca problemฤ de autovaloriโ\] (I-IV, primite la redacลฃia revistei Annalen der Physik รฎn 27 Ianuarie, 23 Februarie, 10 Mai ลi 21 Iunie 1926).
## 6P. Probleme
Problema 6.1 \- Sฤ se obลฃinฤ formulele pentru orbitele stabile ลi pentru nivelele de energie ale electronului รฎn atomul de hidrogen folosind numai argumente bazate pe lungimea de undฤ de Broglie asociatฤelectronului ลi valoarea โempiricฤ โ $`5.310^{11}`$ m pentru raza Bohr.
Soluลฃie: Lungimea de undฤ a electronului este datฤ de $`\lambda =\frac{h}{mv}`$ รฎn timp ce dacฤ egalฤm forลฃa electricฤ cu forลฃa centripetฤ , respectiv $`\frac{mv^2}{r}=\frac{1}{4\pi ฯต_0}\frac{e^2}{r^2}`$ obลฃinem cฤ viteza electronului este datฤ de $`v=\frac{e}{\sqrt{4\pi ฯต_0mr}}.`$รn aceste condiลฃii, lungimea de undฤ a electronului este $`\lambda =\frac{h}{e}\sqrt{\frac{4\pi ฯต_0r}{m}}`$. Acum, dacฤ folosim valoarea $`5.3\times 10^{11}`$m pentru raza $`r`$ a orbitei electronice, vedem cฤ lungimea de undฤ a electronului este $`\lambda =33\times 10^{11}`$ m. Aceastฤ lungime de undฤ are exact aceeaลi valoare ca circumferinลฃa orbitei electronului, $`2\pi r=33\times 10^{11}`$ m. Dupฤ cum se poate vedea, orbita electronului รฎn atomul de hidrogen corespunde astfel unei unde โรฎnchisฤ รฎn ea รฎnsฤลiโ (adicฤ de tip staลฃionar). Acest fapt se poate compara cu vibraลฃiile unui inel de alamฤ . Dacฤ lungimile de undฤ sunt un submultiplu al circumferinลฃei sale, inelul ar putea continua starea sa vibratorie pentru foarte mult timp cu disipare redusฤ (stฤri โpropriiโ de vibraลฃie sau unde staลฃionare). Dacฤ รฎnsฤ numฤrul de lungimi de undฤ nu este รฎntreg se va produce o interferenลฃฤ negativฤ pe mฤsurฤ ce undele se propagฤ de-a lungul inelului ลi vibraลฃiile vor dispฤrea foarte repede. Astfel, se poate afirma cฤ un electron se poate roti indefinit รฎn jurul nucleului fฤrฤ a radia energia de care dispune atรฎta timp cรฎt orbita conลฃine un numฤr รฎntreg de lungimi de undฤ de Broglie. Cu acestea, avem condiลฃia de stabilitate
$`n\lambda =2\pi r_n,`$
unde $`r_n`$ este raza orbitei care conลฃine $`n`$ lungimi de undฤ . Substituind $`\lambda `$, avem
$`{\displaystyle \frac{nh}{e}}\sqrt{{\displaystyle \frac{4\pi ฯต_0r_n}{m}}}=2\pi r_n,`$
ลi deci orbitele stabile ale electronului sunt
$`r_n={\displaystyle \frac{n^2\mathrm{}^2ฯต_0}{\pi me^2}}.`$
Pentru nivelele de energie, avem $`E=T+V`$ ลi prin substituirea energiilor potenลฃialฤ ลi cineticฤ obลฃinem
$`E={\displaystyle \frac{1}{2}}mv^2{\displaystyle \frac{e^2}{4\pi ฯต_0r}},`$
sau echivalent
$`E_n={\displaystyle \frac{e^2}{8\pi ฯต_0r_n}}.`$
Substituind valoarea lui $`r_n`$ รฎn ultima ecuaลฃie obลฃinem
$`E_n={\displaystyle \frac{me^4}{8ฯต_0^2\mathrm{}^2}}\left({\displaystyle \frac{1}{n^2}}\right).`$
Problema 6.2 \- Teorema lui Unsรถld spune cฤ , pentru orice valoare a numฤrului cuantic orbital $`l`$, densitฤลฃile de probabilitate, sumate peste toate substฤrile posibile, de la $`m_l=l`$ pรฎnฤ la $`m_l=+l`$ dau o constantฤ independentฤ de unghiurile $`\theta `$ sau $`\varphi `$, adicฤ
$`{\displaystyle \underset{m_l=l}{\overset{+l}{}}}\mathrm{\Theta }_{lm_l}^2\mathrm{\Phi }_{m_l}^2=ct.`$
Aceastฤ teoremฤ aratฤ cฤ orice atom sau ion cu substฤri รฎnchise prezintฤ o distribuลฃie sferic simetricฤ de sarcinฤ electricฤ . Sฤ se verifice teorema Unsรถld pentru $`l=0`$, $`l=1`$ ลi $`l=2`$.
Soluลฃie: Avem pentru $`l=0`$, $`\mathrm{\Theta }_{00}=1/\sqrt{2}`$ ลi $`\mathrm{\Phi }_0=1/\sqrt{2\pi }`$, deci vedem cฤ
$`\mathrm{\Theta }_{0,0}^2\mathrm{\Phi }_0^2={\displaystyle \frac{1}{4\pi }}.`$
Pentru $`l=1`$ avem
$`{\displaystyle \underset{m_l=1}{\overset{+1}{}}}\mathrm{\Theta }_{lm_l}^2\mathrm{\Phi }_{m_l}^2=\mathrm{\Theta }_{1,1}^2\mathrm{\Phi }_1^2+\mathrm{\Theta }_{1,0}^2\mathrm{\Phi }_0^2+\mathrm{\Theta }_{1,1}^2\mathrm{\Phi }_1^2.`$
Pe de altฤ parte funcลฃiile de undฤ sunt: $`\mathrm{\Theta }_{1,1}=(\sqrt{3}/2)sin\theta `$, $`\mathrm{\Phi }_1=(1/\sqrt{2\pi })e^{i\varphi }`$, $`\mathrm{\Theta }_{1,0}=(\sqrt{6}/2)cos\theta `$, $`\mathrm{\Phi }_0=1/\sqrt{2\pi }`$, $`\mathrm{\Theta }_{1,1}=(\sqrt{3}/2)sin\theta `$, $`\mathrm{\Phi }_1=(1/\sqrt{2\pi })e^{i\varphi }`$ , care substituite รฎn ecuaลฃia anterioarฤ conduc la
$`{\displaystyle \underset{m_l=1}{\overset{+1}{}}}\mathrm{\Theta }_{lm_l}^2\mathrm{\Phi }_{m_l}^2={\displaystyle \frac{3}{8\pi }}sen^2\theta +{\displaystyle \frac{3}{4\pi }}cos^2\theta +{\displaystyle \frac{3}{8\pi }}sen^2\theta ={\displaystyle \frac{3}{4\pi }}`$
ลi din nou obลฃinem o constantฤ .
Pentru $`l=2`$ avem
$`{\displaystyle \underset{m_l=2}{\overset{+2}{}}}\mathrm{\Theta }_{lm_l}^2\mathrm{\Phi }_{m_l}^2=`$
$`\mathrm{\Theta }_{2,2}^2\mathrm{\Phi }_2^2\mathrm{\Theta }_{2,1}^2\mathrm{\Phi }_1^2+\mathrm{\Theta }_{2,0}^2\mathrm{\Phi }_0^2+\mathrm{\Theta }_{2,1}^2\mathrm{\Phi }_1^2+\mathrm{\Theta }_{2,2}^2\mathrm{\Phi }_2^2`$
ลi funcลฃiile de undฤ sunt: $`\mathrm{\Theta }_{2,2}=(\sqrt{15}/4)sin^2\theta `$, $`\mathrm{\Phi }_2=(1/\sqrt{2\pi })e^{2i\varphi }`$, $`\mathrm{\Theta }_{2,1}=(\sqrt{15}/2)sin\theta cos\theta `$, $`\mathrm{\Phi }_1=(1/\sqrt{2\pi })e^{i\varphi }`$, $`\mathrm{\Theta }_{2,0}=(\sqrt{10}/4)(3cos^2\theta 1)`$, $`\mathrm{\Phi }_0=1/\sqrt{2\pi }`$, $`\mathrm{\Theta }_{2,1}=(\sqrt{15}/2)sin\theta cos\theta `$, $`\mathrm{\Phi }_1=(1/\sqrt{2\pi })e^{i\varphi }`$, $`\mathrm{\Theta }_{2,2}=(\sqrt{15}/4)sin^2\theta `$, $`\mathrm{\Phi }_2=(1/\sqrt{2\pi })e^{2i\varphi }`$, care substituite รฎn ecuaลฃia anterioarฤ dau
$`{\displaystyle \underset{m_l=2}{\overset{+2}{}}}\mathrm{\Theta }_{lm_l}^2\mathrm{\Phi }_{m_l}^2={\displaystyle \frac{5}{4\pi }},`$
ceea ce din nou verificฤ teorema Unsรถld.
Problema 6.3 \- Probabilitatea de a gฤsi un electron atomic a cฤrui funcลฃie de undฤ radialฤ este cea de stare fundamentalฤ $`R_{10}(r)`$ รฎn afara unei sfere de razฤ Bohr $`a_0`$ centratฤ รฎn nucleu este
$`{\displaystyle _{a_0}^{\mathrm{}}}R_{10}(r)^2r^2๐r.`$
Sฤ se calculeze probabilitatea de a gฤsi electronul รฎn starea fundamentalฤ atomicฤ la o distanลฃฤ de nucleu mai mare de $`a_0`$.
Soluลฃie: Funcลฃia de undฤ radialฤ care corespunde stฤrii fundamentale este
$`R_{10}(r)={\displaystyle \frac{2}{a_0^{3/2}}}e^{r/a_0}.`$
Substituind-o รฎn integralฤ obลฃinem
$`{\displaystyle _{a_0}^{\mathrm{}}}R(r)^2r^2๐r={\displaystyle \frac{4}{a_0^3}}{\displaystyle _{a_0}^{\mathrm{}}}r^2e^{2r/a_0}๐r,`$
sau
$`{\displaystyle _{a_0}^{\mathrm{}}}R(r)^2r^2๐r={\displaystyle \frac{4}{a_0^3}}\left[{\displaystyle \frac{a_0}{2}}r^2e^{2r/a_0}{\displaystyle \frac{a_0^2}{2}}re^{2r/a_0}{\displaystyle \frac{a_0^3}{4}}e^{2r/a_0}\right]_{a_0}^{\mathrm{}}.`$
Aceasta ne conduce la:
$`{\displaystyle _{a_0}^{\mathrm{}}}R(r)^2r^2๐r={\displaystyle \frac{5}{e^2}}68\%!!,`$
care este probabilitatea cerutฤ รฎn aceastฤ problemฤ .
7. CIOCNIRI CUANTICE
## Introducere
Pentru iniลฃiere รฎn teoria cuanticฤ de รฎmprฤลtiere ne vom servi de rezultate deja cunoscute de la รฎmprฤลtierea clasicฤ รฎn cรฎmpuri centrale ลi vom presupune anumite situaลฃii care vor simplifica calculele fฤrฤ รฎnsฤ a ne รฎndepฤrta prea mult de problema โrealฤโ. ลtim cฤ รฎn studiul experimental al unei ciocniri putem obลฃine date care ne pot ajuta sฤ รฎnลฃelegem distribuลฃia materiei โลฃintฤโ, sau mai bine spus interacลฃiunea รฎntre fasciculul incident ลi โลฃintฤโ. Ipotezele pe care le vom presupune corecte sunt:
i) Particulele nu au spin, ceea ce nu รฎnseamnฤ cฤ acesta nu este important รฎn ciocniri.
ii) Ne vom ocupa numai de dispersia elasticฤ pentru care posibila structurฤ internฤ a particulelor nu se ia รฎn considerare.
iii) ลขinta este suficient de subลฃire pentru a putea neglija รฎmprฤลtierile multiple.
iv) Interacลฃiunile sunt descrise printr-un potenลฃial care depinde numai de poziลฃia relativฤ a particulelor.
Aceste ipoteze eliminฤ o serie de efecte cuantice ลi mฤ resc corectitudinea unor rezultate bazice din teoria ciocnirilor clasice. Astfel definim:
$$\frac{d\sigma }{d\mathrm{\Omega }}=\frac{I(\theta ,\phi )}{I_0},$$
(1)
unde $`d\sigma `$ este elementul de unghi solid, $`I_0`$ este numฤrul de particule incidente pe unitate de arie ลi $`Id\mathrm{\Omega }`$ este numฤrul de particule dispersate รฎn elementul de unghi solid $`d\mathrm{\Omega }`$.
Cu aceste concepte binecunoscute ลi cu ajutorul mฤrimii asimptotice parametru de impact $`b`$ asociat fiecฤrei particule clasice incidente ajungem la importanta formulฤ clasicฤ
$$\frac{d\sigma }{d\mathrm{\Omega }}=\frac{b}{\mathrm{sin}\theta }|\frac{db}{d\theta }|.$$
(2)
Dacฤ dorim sฤ cunoaลtem รฎn termeni cuantici fenomenologia de ciocnire, trebuie sฤ studiem evoluลฃia รฎn timp a unui pachet de unde. Fie $`F_i`$ fluxul de particule al fascicolului incident, adicฤ , numฤrul de particule pe unitate de timp care intersecteazฤ o suprafaลฃฤ unitarฤ transversalฤ axei de propagare. Vom poziลฃiona un detector departe de zona de acลฃiune efectivฤ a potenลฃialului, care subรฎntinde un unghi solid $`d\mathrm{\Omega }`$; cu acesta putem รฎnregistra numฤrul de particule $`dn/dt`$ dispersate รฎn unitatea de timp รฎn $`d\mathrm{\Omega }`$ รฎn direcลฃia $`(\theta ,\phi )`$.
$`dn/dt`$ este proporลฃional cu $`d\mathrm{\Omega }`$ ลi $`F_i`$. Sฤ numim $`\sigma (\theta ,\phi )`$ coeficientul de proporลฃionalitate รฎntre $`dn`$ ลi $`F_id\mathrm{\Omega }`$:
$$dn=\sigma (\theta ,\phi )F_id\mathrm{\Omega },$$
(3)
care este prin definiลฃie secลฃiunea diferenลฃialฤ transversalฤ .
Numฤrul de particule pe unitatea de timp care ajung la detector este egal cu numฤrul de particule care intersecteazฤ o suprafaลฃฤ $`\sigma (\theta ,\phi )d\mathrm{\Omega }`$ situatฤ perpendicular pe axa fasciculului. Secลฃiunea totalฤ de dispersie este prin definiลฃie:
$$\sigma =\sigma (\theta ,\phi )๐\mathrm{\Omega }.$$
(4)
Cum putem orienta axele de coordonate conform alegerii dorite, o vom face รฎn aลa fel ca axa fasciculului incident de particule sฤ coincidฤ cu axa z (aceasta pentru simplificarea calculelor, unde vom folosi coordonatele sferice).
รn regiunea negativฤ a axei, pentru $`t`$ negativ mare, particula va fi practic liberฤ : nu este afectatฤ de $`V(๐ซ)`$ ลi starea sa se poate reprezenta prin unde plane. Prin urmare funcลฃia de undฤ trebuie sฤ conลฃinฤ termeni de forma $`e^{ikz}`$, unde $`k`$ este constanta care apare รฎn ecuaลฃia Helmholtz. Prin analogie cu optica, forma undei dispersate este:
$$f(r)=\frac{e^{ikr}}{r}.$$
(5)
รntr-adevฤr:
$$(^2+k^2)e^{ikr}0$$
(6)
ลi
$$(^2+k^2)\frac{e^{ikr}}{r}=0$$
(7)
pentru $`r>r_0`$, unde $`r_0`$ este orice numฤr pozitiv.
Presupunem cฤ miลcarea particulei este descrisฤ de Hamiltonianul:
$$H=\frac{๐ฉ^\mathrm{๐}}{2\mu }+V=H_0+V.$$
(8)
V este diferit de zero numai รฎntr-o micฤ vecinฤtate รฎn jurul originii. ลtim cฤ un pachet de unde รฎn $`t=0`$ se poate scrie:
$$\psi (๐ซ,0)=\frac{1}{(2\pi )^{\frac{3}{2}}}\phi (๐ค)\mathrm{exp}[i๐ค(๐ซ๐ซ_\mathrm{๐})]๐^\mathrm{๐}๐ค,$$
(9)
unde $`\psi `$ este o funcลฃie semnificativ nenulฤ รฎn segmentul (lฤrgimea) $`\mathrm{\Delta }๐ค`$ centrat รฎn jurul lui $`๐ค_\mathrm{๐}`$. Presupunem deasemena cฤ $`๐ค_\mathrm{๐}`$ este paralel la $`๐ซ_\mathrm{๐}`$, dar de sens opus. Pentru a vedea รฎn mod cantitativ ce se รฎntรฎmplฤ cu pachetul de unde cรฎnd la un moment ulterior ciocneลe ลฃinta ลi este dispersat de aceasta ne putem folosi de dezvoltarea lui $`\psi (๐ซ,0)`$ รฎn funcลฃiile proprii $`\psi _n(๐ซ)`$ ale lui $`H`$, respectiv $`\psi (๐ซ,0)=_nc_n\psi _n(๐ซ)`$. Astfel, pachetul de unde la timpul $`t`$ este:
$$\psi (๐ซ,t)=\underset{n}{}c_n\phi _n(๐ซ)\mathrm{exp}(\frac{i}{\mathrm{}}E_nt).$$
(10)
Aceasta este o funcลฃie proprie a operatorului $`H_0`$ ลi nu a lui $`H`$, dar putem substitui aceste funcลฃii proprii cu funcลฃii proprii particulare ale lui $`H`$, pe care le vom nota cu $`\psi _k^{(+)}(๐ซ)`$. Forma asimptoticฤ a acestora din urmฤ este de tipul:
$$\psi _k^{(+)}(๐ซ)๐^{\mathrm{๐ข๐ค}๐ซ}+๐(๐ซ)\frac{๐^{\mathrm{๐ข๐ค๐ซ}}}{๐ซ},$$
(11)
unde, cum este uzual, $`๐ฉ=\mathrm{}๐ค`$ ลi $`E=\frac{\mathrm{}^2k^2}{2m}`$.
Aceasta corespunde unei unde plane ca fascicul incident ลi o undฤ sfericฤ divergentฤ, despre care se poate spune cฤ este rezultatul interacลฃiunii รฎntre fascicul ลi ลฃintฤ. Aceste soluลฃii ale ec. Schrรถdinger existฤ รฎn realitate, ลi putem dezvolta $`\psi (๐ซ,0)`$ รฎn unde plane ลi $`\psi _k(๐ซ)`$:
$$\psi (๐ซ,0)=\phi (๐ค)\mathrm{exp}(i๐ค๐ซ_\mathrm{๐})\psi _๐ค(๐ซ)d^3k,$$
(12)
unde $`\mathrm{}\omega =\frac{\mathrm{}^2k^2}{2m}`$. Se poate spune deci cฤ unda sfericฤ divergentฤ nu are nici o contribuลฃie la pachetul de unde iniลฃial.
## รmprฤลtierea unui pachet de unde
Orice undฤ suferฤ รฎn cursul propagฤrii o dispersie. De aceea nu se poate ignora efectul undei divergente din acest punct de vedere. Se poate folosi urmฤtorul truc:
$$\omega =\frac{\mathrm{}}{2m}k^2=\frac{\mathrm{}}{2m}[๐ค_\mathrm{๐}+(๐ค๐ค_\mathrm{๐})]^2=\frac{\mathrm{}}{2m}[2๐ค_\mathrm{๐}๐ค๐ค_\mathrm{๐}^\mathrm{๐}+(๐ค๐ค_\mathrm{๐})^\mathrm{๐}],$$
(13)
pentru a neglija ultimul termen รฎn paranteze. Substituind $`\omega `$ รฎn $`\psi `$, cerem ca: $`\frac{\mathrm{}}{2m}(๐ค๐ค_\mathrm{๐})^2T1`$, unde $`T\frac{2mr_0}{\mathrm{}k_0}`$ ลi deci:
$$\frac{(\mathrm{\Delta }k)^2r_0}{k_0}1.$$
(14)
Aceastฤ condiลฃie ne spune cฤ pachetul de unde nu se disperseazฤ รฎn mod apreciabil chiar ลi atunci cรฎnd se deplaseazฤ pe o distanลฃฤ macroscopicฤ $`r_0`$.
Alegรฎnd direcลฃia vectorului $`๐ค`$ al undei incidente de-a lungul uneia dintre cele trei direcลฃii carteziene, putem scrie รฎn coordonate sferice
$`\psi _k(r,\theta ,\phi )e^{ikz}+\frac{f(k,\theta ,\phi )e^{ikr}}{r}.`$
รntrucรฎt $`H`$, operatorul Hamiltonian (que hemos considerat pรฎnฤ acum nu ca operator pentru cฤ rezultatele sunt aceleaลi) este invariant la rotaลฃiile รฎn axa z, putem alege condiลฃiile de frontierฤ deasemenea invariante, astfel cฤ :
$`\psi _k(r,\theta ,\phi )e^{ikz}+\frac{f(\theta )e^{ikr}}{r}.`$
Acest tip de funcลฃii se cunosc ca unde de รฎmprฤลtiere. Coeficientul $`f(\theta )`$ al undei divergente se cunoaลte ca amplitudine de รฎmprฤลtiere.
## Amplitudinea de probabilitate รฎn รฎmprฤลtieri
Ecuaลฃia Schrรถdinger de rezolvat este:
$$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2m}^2\psi +V(๐ซ,t)\psi .$$
(15)
Expresia
$$P(๐ซ,t)=\psi ^{}(๐ซ,t)\psi (๐ซ,t)=|\psi (๐ซ,t)|^2$$
(16)
se interpreteazฤ , cf. lui Max Born, ca o densitate de probabilitate dacฤ funcลฃia de undฤ se normalizeazฤ astfel ca:
$$|\psi (๐ซ,t)|^2d^3r=1.$$
(17)
Desigur integrala de normalizare a lui $`\psi `$ trebuie sฤ fie independentฤ de timp. Acest lucru se poate nota รฎn felul urmฤtor:
$$I=\frac{}{t}_\mathrm{\Omega }P(๐ซ,t)d^3r=_\mathrm{\Omega }(\psi ^{}\frac{\psi }{t}+\frac{\psi ^{}}{t}\psi )d^3r$$
(18)
ลi din ec. Schrรถdinger:
$$\frac{\psi }{t}=\frac{i\mathrm{}}{2m}^2\psi \frac{i}{\mathrm{}}V(๐ซ,t)\psi $$
(19)
rezultฤ :
$$I=\frac{i\mathrm{}}{2m}_\mathrm{\Omega }[\psi ^{}^2(^2\psi ^{})\psi ]d^3r=\frac{i\mathrm{}}{2m}_\mathrm{\Omega }[\psi ^{}\psi (\psi ^{})\psi ]d^3r=$$
$$=\frac{i\mathrm{}}{2m}_A[\psi ^{}\psi (\psi ^{})\psi ]_n๐A,$$
(20)
unde s-a folosit teorema Green pentru evaluarea integralei de volum. $`dA`$ este elementul de suprafaลฃฤ pe frontiera care delimiteazฤ regiunea de integrare ลi $`[]_n`$ denotฤ componenta รฎn direcลฃia normalฤ la elementul de suprafaลฃฤ $`dA`$.
Definind:
$$๐(๐ซ,t)=\frac{\mathrm{}}{2im}[\psi ^{}\psi (\psi ^{})\psi ],$$
(21)
obลฃinem:
$$I=\frac{}{t}_\mathrm{\Omega }P(๐ซ,t)d^3r=_\mathrm{\Omega }๐d^3r=_AS_n๐A,$$
(22)
pentru pachete de undฤ รฎn care $`\psi `$ se pune zero la distanลฃe mari ลi integrala de normalizare converge, integrala de suprafaลฃฤ este zero cรฎnd $`\mathrm{\Omega }`$ este tot spaลฃiul. Se poate demonstra (se poate consulta P. Dennery & A. Krzywicki, Mathematical methods for physicists) cฤ integrala de suprafaลฃฤ este zero, astfel cฤ integrala de normalizare este constantฤ รฎn timp ลi deci se satisface cerinลฃa iniลฃialฤ . Din aceeaลi ecuaลฃie pentru $`๐`$ obลฃinem:
$$\frac{P(๐ซ,t)}{t}+๐(๐ซ,t)=0,$$
(23)
care este o ecuaลฃie de continuitate cu fluxul de densitate $`P`$ ลi curent de densitate $`๐`$, fฤrฤ nici un fel de surse (pozitive sau negative). Dacฤ interpretฤm $`\frac{\mathrm{}}{im}`$ ca un fel de โoperatorโ vitezฤ (ca ลi รฎn cazul timpului nu se poate vorbi de un operator vitezฤ รฎn sens riguros), atunci:
$$๐(๐ซ,t)=Re(\psi ^{}\frac{\mathrm{}}{im}\psi ).$$
(24)
Calculul efectiv al densitฤลฃii de curent pentru o undฤ de รฎmprฤลtiere este de tip โtrucโ ลi nu รฎl considerฤm ilustrativ. Rezultatul final este $`j_r=\frac{hk}{mr^2}|f(\theta )|^2`$, unde direcลฃia $`\theta =0`$ nu se include.
## Funcลฃia Green รฎn teoria de รฎmprฤลtiere
O altฤ formฤ de a scrie ecuaลฃia Schrรถdinger de rezolvat este $`(\frac{\mathrm{}^2}{2m}^2+V)\psi =E\psi `$ sau $`(^2+k^2)\psi =U\psi `$ unde: $`k^2=\frac{2mE}{\mathrm{}^2}`$ ลi $`U=\frac{2mV}{\mathrm{}^2}`$.
Rezultฤ mai convenabil de transformat aceastฤ ecuaลฃie la o formฤ integralฤ . Aceasta se poate face dacฤ vom considera $`U\psi `$ din partea dreaptฤ a ecuaลฃiei ca o inomogeneitate, ceea ce ne permite sฤ construim soluลฃia ecuaลฃiei cu ajutorul funcลฃiei Green (nucleu integral), care prin definiลฃie este soluลฃia lui:
$$(^2+k^2)G(๐ซ,๐ซ^{})=\delta (๐ซ๐ซ^{}).$$
(25)
Soluลฃia ecuaลฃiei Schrรถdinger se dฤ ca suma soluลฃiei ecuaลฃiei omogene ลi a soluลฃiei inomogene de tip Green:
$$\psi (๐ซ)=\lambda (๐ซ)๐(๐ซ,๐ซ^{})๐(๐ซ^{})\psi (๐ซ^{})๐^\mathrm{๐}๐ซ^{}.$$
(26)
Cฤutฤm acum o funcลฃie $`G`$ care sฤ fie un produs de funลฃii linear independente, cum sunt de exemplu undele plane:
$$G(๐ซ,๐ซ^{}=A(๐ช)e^{i๐ช(๐ซ๐ซ^{})}dq.$$
(27)
Folosind ecuaลฃia 25, avem:
$$A(๐ช)(k^2q^2)e^{i๐ช(๐ซ๐ซ^{})}๐q=\delta (๐ซ๐ซ^{}),$$
(28)
care se transformฤ รฎntr-o identitate dacฤ :
$$A(๐ช)=(2\pi )^3(k^2q^2)^1.$$
(29)
De aici rezultฤ :
$$G(๐ซ,๐ซ^{})=\frac{1}{(2\pi )^3}\frac{e^{iqR}}{k^2q^2}d^3q,$$
(30)
cu $`R=|๐ซ๐ซ^{}|`$. Dupฤ un calcul folosind metode de variabilฤ complexฤ <sup>10</sup><sup>10</sup>10Se poate vedea problema 7.1., ajungem la:
$$G(R)=\frac{1}{4\pi }\frac{e^{ikR}}{R}.$$
(31)
Aceastฤ funcลฃie nu este determinatฤ รฎn mod univoc; funcลฃia Green poate fi oricare soluลฃie a ecuaลฃiei 25; Alegerea uneia particulare se face prin impunerea condiลฃiilor de frontierฤ asupra funcลฃiilor proprii $`\psi _k(๐ซ)`$.
Funcลฃia Green obลฃinutฤ รฎn aceste condiลฃii este:
$$G(๐ซ,๐ซ^{})=\left(\frac{e^{ik|๐ซ๐ซ^{}|}}{4\pi |๐ซ๐ซ^{}|}\right).$$
(32)
รn acest fel, ajungem la ecuaลฃia integralฤ pentru funcลฃia de undฤ de ciocnire:
$$\psi (k,๐ซ)=\phi (k,๐ซ)\frac{m}{2\pi \mathrm{}^2}\frac{e^{ik|๐ซ๐ซ^{}|}}{๐ซ๐ซ^{}}U(๐ซ^{})\psi (k,๐ซ)๐๐ซ,$$
(33)
unde $`\phi `$ este o soluลฃie a ecuaลฃiei Helmholtz. Notรฎnd $`|๐ซ๐ซ^{}|=R`$:
$$(^2+k^2)\psi =(^2+k^2)[\phi +G(๐ซ,๐ซ^{})U(๐ซ^{})\psi (๐ซ^{})d^3r^{}]$$
(34)
ลi presupunรฎnd cฤ putem schimba ordinea operaลฃiilor ลi pune operatorul $`^2`$ รฎn interiorul integralei:
$$(^2+k^2)\psi =(^2+k^2)G(๐ซ,๐ซ^{})U(๐ซ^{})\psi (๐ซ^{})d^3r^{}=U(๐ซ)\psi (๐ซ),$$
(35)
ceea ce ne aratฤ cฤ se verificฤ faptul cฤ $`G(R)=\frac{1}{4\pi }\frac{e^{ikR}}{R}`$ este soluลฃie.
## Teorema opticฤ
Secลฃiunea diferenลฃialฤ totalฤ este datฤ de:
$$\sigma _{tot}(k)=\frac{d\sigma }{d\mathrm{\Omega }}๐\mathrm{\Omega }.$$
(36)
Sฤ exprimฤm acum $`f(\theta )`$ ca funcลฃie de ลiftul de fazฤ $`S_l(k)=e^{2i\delta _l(k)}`$ รฎn forma:
$$f(\theta )=\frac{1}{k}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)e^{i\delta _i(k)}\mathrm{sin}\delta _l(k)P_l(\mathrm{cos}\theta )$$
(37)
atunci
$$\sigma _{tot}=[\frac{1}{k}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)e^{i\delta _l(k)}\mathrm{sin}\delta _l(k)P_l(\mathrm{cos}\theta )]$$
$$[[\frac{1}{k}\underset{l^{}=0}{\overset{\mathrm{}}{}}(2l^{}+1)e^{i\delta _l^{}(k)}\mathrm{sin}\delta _l^{}(k)P_l^{}(\mathrm{cos}\theta )].$$
(38)
Folosind acum $`P_l(\mathrm{cos}\theta )P_l^{}(\mathrm{cos}\theta )=\frac{4\pi }{2l+1}\delta _{ll^{}}`$ obลฃinem
$$\sigma _{tot}=\frac{4\pi }{k^2}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)\mathrm{sin}\delta _l(k)^2.$$
(39)
Ceea ce ne intereseazฤ este cฤ :
$$\mathrm{Im}f(0)=\frac{1}{k}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)\mathrm{Im}[e^{i\delta _l(k)}\mathrm{sin}\delta _l(k)]P_l(1)=\frac{1}{k}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)\mathrm{sin}\delta _l(k)^2=$$
$$\frac{k}{4\pi }\sigma _{tot}.$$
(40)
Aceastฤ relaลฃie este cunoscutฤ ca teorema opticฤ . Semnificaลฃia sa fizicฤ este cฤ interferenลฃa undei incidente cu unda dispersatฤ la unghi zero produce โieลireaโ particulei din unda incidentฤ , ceea ce permite conservarea probabilitฤลฃii.
## Aproximaลฃia Born
Sฤ considerฤm situaลฃia din Fig. 7.2:
Punctul de observare M este departe de P, care se aflฤ รฎn regiunea de influenลฃฤ a potenลฃialului $`U`$, cu $`rL`$, $`r^{}l`$. Segmentul MP, care corespunde la $`|๐ซ๐ซ^{}|`$, este รฎn aceste condiลฃii geometrice aproximativ egal cu proiecลฃia lui MP pe MO:
$$|๐ซ๐ซ^{}|r๐ฎ๐ซ^{},$$
(41)
unde $`๐ฎ`$ este vectorul unitar (versor) รฎn direcลฃia $`๐ซ`$. Atunci, pentru $`r`$ mare:
$$G=\frac{1}{4\pi }\frac{e^{ik|๐ซ๐ซ^{}|}}{|๐ซ๐ซ^{}|}_r\mathrm{}\frac{1}{4\pi }\frac{e^{ikr}}{r}e^{ik๐ฎ๐ซ}.$$
(42)
Substituim $`G`$ รฎn expresia integralฤ a funcลฃiei de undฤ de ciocnire pentru a obลฃine:
$$\psi (๐ซ)=e^{ikz}\frac{1}{4\pi }\frac{e^{ikr}}{r}e^{ik๐ฎ๐ซ}U(๐ซ^{})\psi (๐ซ^{})d^3r^{}.$$
(43)
Aceasta deja nu mai este o funcลฃie de distanลฃa $`r=OM`$, ci numai de $`\theta `$ ลi $`\psi `$; atunci:
$$f(\theta ,\psi )=\frac{1}{4\pi }e^{ik๐ฎ๐ซ}U(๐ซ^{})\psi (๐ซ^{})d^3r^{}.$$
(44)
Definim acum vectorul de undฤ incident $`๐ค_๐ข`$ ca un vector de modul $`k`$ dirijat de-a lungul axei polare a fasciculului astfel cฤ : $`e^{ikz}=e^{i๐ค_๐ข๐ซ}`$; similar, $`๐ค_๐`$, de modul $`k`$ ลi cu direcลฃia fixatฤ prin $`\theta `$ ลi $`\phi `$, se numeลte vector de undฤ โdeplasatโ รฎn direcลฃia $`(\theta ,\phi )`$: $`๐ค_๐=k๐ฎ`$
Vectorul de undฤ transferat รฎn direcลฃia $`(\theta ,\phi )`$ se introduce prin: $`๐=๐ค_๐๐ค_๐ข`$.
Cu aceasta, putem scrie ecuaลฃia integralฤ de dispersie รฎn forma:
$$\psi (๐ซ)=e^{i๐ค_๐ข๐ซ}+G(๐ซ,๐ซ^{})U(๐ซ^{})\psi (๐ซ^{})d^3r^{}$$
(45)
Acum putem รฎncerca rezolvarea acestei ecuaลฃii รฎn mod iterativ. Punรฎnd $`๐ซ๐ซ^{};๐ซ^{}๐ซ^{\prime \prime }`$, putem scrie:
$$\psi (๐ซ^{})=e^{i๐ค_๐ข๐ซ^{}}+G(๐ซ^{},๐ซ^{\prime \prime })U(๐ซ^{\prime \prime })\psi (๐ซ^{\prime \prime })d^3r^{\prime \prime }.$$
(46)
Substituind รฎn 45 obลฃinem:
$$\psi (๐ซ)=e^{i๐ค_ir}+G(๐ซ,๐ซ^{})U(๐ซ^{})e^{i๐ค_๐ข๐ซ^{}}d^3r^{}+$$
$$G(๐ซ,๐ซ^{})U(๐ซ^{})G(๐ซ^{},๐ซ^{\prime \prime })U(๐ซ^{\prime \prime })\psi (๐ซ^{\prime \prime })d^3r^{\prime \prime }d^3r^{}.$$
(47)
Primii doi termeni din partea dreaptฤ sunt cunoscuลฃi ลi numai al treilea conลฃine funcลฃia necunoscutฤ $`\psi (๐ซ)`$. Putem repeta acest procedeu: รฎnlocuind $`๐ซ`$ cu $`๐ซ^{\prime \prime }`$ ลi $`๐ซ^{}`$ cu $`๐ซ^{\prime \prime \prime }`$ obลฃinem $`\psi (๐ซ^{\prime \prime })`$ , pe care putem sฤ o reintroducem รฎn ec. 47:
$$\psi (๐ซ)=e^{i๐ค_๐ข๐ซ}+G(๐ซ,๐ซ^{})U(๐ซ^{})e^{i๐ค_๐ข๐ซ^{}}+$$
$$G(๐ซ,๐ซ^{})U(๐ซ^{})G(๐ซ^{},๐ซ^{\prime \prime })U(๐ซ^{\prime \prime })e^{i๐ค_๐ข๐ซ^{\prime \prime }}d^3r^{}d^3r^{\prime \prime }+$$
$$G(๐ซ,๐ซ^{})U(๐ซ^{})G(๐ซ^{},๐ซ^{\prime \prime })U(๐ซ^{\prime \prime })e^{i๐ค_๐ข๐ซ^{\prime \prime }}G(๐ซ^{\prime \prime },๐ซ^{\prime \prime \prime })U(๐ซ^{\prime \prime \prime })\psi (๐ซ^{\prime \prime \prime }).$$
(48)
Primii trei termeni sunt cunoscuลฃi; funcลฃia necunoscutฤ $`\psi (๐ซ)`$ se aflฤ รฎn al patrulea termen. รn acest fel, prin iteraลฃii construim funcลฃia de undฤ de dispersie staลฃionarฤ . Notฤm cฤ fiecare termen รฎn dezvoltarea รฎn serie prezintฤ o putere superioarฤ รฎn potenลฃial faลฃฤ de cel precedent. Putem continua รฎn acest fel pรฎnฤ cรฎnd obลฃinem o expresie neglijabilฤ รฎn partea dreaptฤ , ลi obลฃinem $`\psi (๐ซ)`$ รฎn funcลฃie numai de mฤrimi cunoscute.
Substituind expresia lui $`\psi (๐ซ)`$ รฎn $`f(\theta ,\phi )`$ obลฃinem dezvoltarea รฎn serie Born a amplitudinii de รฎmprฤลtiere. Limitรฎndu-ne la primul ordin รฎn $`U`$, trebuie sฤ se facฤ doar substituirea lui $`\psi (๐ซ^{})`$ cu $`e^{i๐ค_๐ข๐ซ^{}}`$ รฎn partea dreaptฤ a ecuaลฃiei pentru a obลฃine:
$$f^{(B)}(\theta ,\phi )=\frac{1}{4\pi }e^{i๐ค_๐ข๐ซ^{}}U(๐ซ^{})e^{ik๐ฎ๐ซ^{}}d^3r^{}=\frac{1}{4\pi }e^{i(๐ค_๐๐ค_๐ข)๐ซ^{}}U(๐ซ^{})d^3r^{}=$$
$$\frac{1}{4\pi }e^{i๐๐ซ^{}}U(๐ซ^{})d^3r^{}$$
(49)
$`๐`$ este vectorul de undฤ transferat definit mai รฎnaite. Vedem cฤ secลฃiunea de dispersie se relaลฃioneazฤ รฎn mod simplu cu potenลฃialul, dacฤ ลฃinem cont de $`V(๐ซ)=\frac{\mathrm{}^2}{2m}U(๐ซ)`$ ลi $`\sigma (\theta ,\phi )=|f(\theta ,\phi )|^2`$. Rezultatul este:
$$\sigma ^{(B)}(\theta ,\phi )=\frac{m^2}{4\pi ^2\mathrm{}^4}|e^{i๐๐ซ}V(๐ซ)d^3r|^2$$
(50)
Direcลฃia ลi modulul vectorului undei dispersate $`๐`$ depinde de modulul $`k`$ al lui $`๐ค_๐ข`$ ลi $`๐ค_๐`$ precum ลi de direcลฃia de รฎmprฤลtiere $`(\theta ,\phi )`$. Pentru $`\theta `$ ลi $`\phi `$ daลฃi, secลฃiunea eficace este o funcลฃie de $`k`$, energia fasciculului incident. Analog, pentru o energie datฤ , $`\sigma ^{(B)}`$ este o funcลฃie de $`\theta `$ ลi $`\phi `$. Aproximaลฃia Born permite ca studiind variaลฃia secลฃiunii eficace diferenลฃiale รฎn funcลฃie de direcลฃia de รฎmprฤลtiere ลi energia incidentฤ sฤ obลฃinem informaลฃii asupra potenลฃialului $`V(๐ซ)`$.
7N. Notฤ: Unul dintre primele articole de รฎmprฤลtiere cuanticฤ este:
M. Born, โQuantenmechanik der Stossvorgรคngeโ \[โMecanica cuanticฤ a proceselor de ciocnireโ\], Zf. f. Physik 37, 863-867 (1926)
## 7P. Probleme
Problema 7.1
Calculul de variabilฤ complexฤ a funcลฃiei Green
Reamintim cฤ am obลฃinut deja rezultatul:
$`G(๐ซ,๐ซ^{})=\frac{1}{(2\pi )^3}\frac{e^{iqR}}{k^2q^2}d^3q,`$ cu $`R=|๐ซ๐ซ^{}|`$. Cum $`d^3q=q^2\mathrm{sin}\theta dqd\theta d\varphi `$, ajungem, dupฤ ce integrฤm รฎn variabilele unghiulare , la:
$`G(๐ซ,๐ซ^{})=\frac{i}{4\pi ^2R}_{\mathrm{}}^{\mathrm{}}\frac{(e^{iqR}e^{iqR})}{k^2q^2}q๐q.`$
Sฤ punem: $`C=\frac{i}{4\pi ^2R}`$; ลi sฤ separฤm integrala รฎn douฤ pฤrลฃi:
$`C(_{\mathrm{}}^{\mathrm{}}\frac{e^{iqR}}{k^2q^2}q๐q_{\mathrm{}}^{\mathrm{}}\frac{e^{iqR}}{k^2q^2}q๐q).`$
Sฤ facem acum $`qq`$ รฎn prima integralฤ :
$`_{\mathrm{}}^{\mathrm{}}\frac{e^{i(q)R}}{k^2(q)^2}(q)d(q)=_{\mathrm{}}^{\mathrm{}}\frac{e^{iqR}}{k^2q^2}q๐q=_{\mathrm{}}^{\mathrm{}}\frac{e^{iqR}}{k^2q^2}q๐q`$
astfel cฤ :
$`G(๐ซ,๐ซ^{})=2C(_{\mathrm{}}^{\mathrm{}}\frac{qe^{iqR}}{k^2q^2}๐q).`$
Substituind $`C`$, obลฃinem:
$`G(๐ซ,๐ซ^{})=\frac{i}{2\pi ^2R}_{\mathrm{}}^{\mathrm{}}\frac{qe^{iqR}}{k^2q^2}๐q`$
รn aceastฤ formฤ integrala se poate evalua cu ajutorul reziduurilor polilor pe care รฎi posedฤ , folosind metodele de variabilฤ complexฤ . Notฤm cฤ existฤ poli simpli รฎn poziลฃiile $`q=_{}^+k`$.
Fig. 7.4: Reguli de contur รฎn jurul polilor pentru $`G_+`$ ลi $`G_{}`$
Folosim conturul din figura 7.4, care รฎnconjoarฤ polii รฎn modul arฤtat, pentru cฤ acesta dฤ efectul fizic corect, pentru cฤ de acord cu teorema reziduurilor,
$`G(r)=\frac{1}{4\pi }\frac{e^{ikr}}{r}(\mathrm{Im}k>0)`$ ,
$`G(r)=\frac{1}{4\pi }\frac{e^{ikr}}{r}(\mathrm{Im}k<0)`$
Soluลฃia care ne intereseazฤ este prima, pentru cฤ dฤ unde dispersate divergente, รฎn timp ce a doua soluลฃie reprezintฤ unde dispersate convergente. Mai mult, combinaลฃia linearฤ
$`\frac{1}{2}lim_{ฯต0}[G_{k+iฯต}+G_{kiฯต}]=\frac{1}{4\pi }\frac{\mathrm{cos}kr}{r}`$
corespunde undelor staลฃionare.
Evaluarea formalฤ a integralei se poate face luรฎnd $`k^2q^2k^2+iฯตq^2`$ , astfel cฤ : $`_{\mathrm{}}^{\mathrm{}}\frac{qe^{iqR}}{k^2q^2}๐q_{\mathrm{}}^{\mathrm{}}\frac{qe^{iqR}}{(k^2+iฯต)q^2}๐q.`$
Aceasta este posibil pentru $`R>0`$, de aceea conturul pentru calcul va fi situat รฎn semiplanul complex superior. Astfel, polii integrantului se aflฤ รฎn: $`q=_{}^+\sqrt{k^2+iฯต}_{}^+(k+\frac{iฯต}{2k})`$. Procedeul de luare a limitei cรฎnd $`ฯต0`$ trebuie efectuat dupฤ evaluarea integralei.
Problema 7.2
Forma asimptoticฤ a funcลฃiei radiale
Cum s-a vฤzut deja รฎn capitolul Atomul de hidrogen partea radialฤ a ec. Schrรถdinger se poate scrie:
$`(\frac{d^2}{dr^2}+\frac{2}{r}\frac{d}{dr})R_{nlm}(r)\frac{2m}{\mathrm{}^2}[V(r)+\frac{l(l+1)\mathrm{}^2}{2mr^2}]R_{nlm}(r)+\frac{2mE}{\mathrm{}^2}R_{nlm}(r)=0.`$
$`n,l,m`$ sunt numerele cuantice sferice. De acum รฎnainte nu se vor mai scrie din motive de comoditate. $`R`$ este funcลฃia de undฤ radialฤ (depinde numai de $`r`$). Vom presupune cฤ potenลฃialele cad la zero mai repede decรฎt $`1/r`$, ลi รฎn plus cฤ $`lim_{r0}r^2V(r)=0`$.
Folosim acum $`u(r)=rR`$, ลi cum: $`(\frac{d^2}{dr^2}+\frac{2}{r}\frac{d}{dr})\frac{u}{r}=\frac{1}{r}\frac{d^2}{dr^2}u`$, avem
$`\frac{d^2}{dr^2}u+\frac{2m}{\mathrm{}^2}[EV(r)\frac{l(l+1)\mathrm{}^2}{2mr^2}]u=0.`$
Notฤm cฤ potenลฃialul prezintฤ un termen suplimentar:
$`V(r)V(r)+\frac{l(l+1)\mathrm{}^2}{2mr^2},`$
care corespunde unei bariere centrifugale repulsive. Pentru o particulฤ liberฤ $`V(r)=0`$ ลi ecuaลฃia devine
$`[\frac{d^2}{dr^2}+\frac{2}{r}\frac{d}{dr})\frac{l(l+1)}{r^2}]R+k^2R=0.`$
Introducรฎnd variabila $`\rho =kr`$, obลฃinem
$`\frac{d^2R}{d\rho ^2}+\frac{2}{\rho }\frac{dR}{d\rho }\frac{l(l+1)}{\rho ^2}R+R=0.`$
Soluลฃiile acestei ecuaลฃii sunt aลa numitele funcลฃii Bessel sferice. Soluลฃia regularฤ este:
$`j_l(\rho )=(\rho )^l(\frac{1}{\rho }\frac{d}{d\rho })^l(\frac{\mathrm{sin}\rho }{\rho }),`$
iar cea iregularฤ :
$`n_l(\rho )=(\rho )^l(\frac{1}{\rho }\frac{d}{d\rho })^l(\frac{\mathrm{cos}\rho }{\rho }).`$
Pentru $`\rho `$ mare, funcลฃiile de interes sunt funcลฃiile Hankel sferice:
$`h_l^{(1)}(\rho )=j_l(\rho )+in_l(\rho )`$ ลi $`h_l^{(2)}(\rho )=[h_l^{(1)}(\rho )]^{}.`$
De interes deosebit este comportamentul pentru $`\rho l`$:
$$j_l(\rho )\frac{1}{\rho }\mathrm{sin}(\rho \frac{l\pi }{2})$$
(51)
$$n_l(\rho )\frac{1}{\rho }\mathrm{cos}(\rho \frac{l\pi }{2}).$$
(52)
ลi atunci
$`h_l^{(1)}\frac{i}{\rho }e^{i(\rho l\pi /2)}.`$
Soluลฃia regularฤ รฎn origine este: $`R_l(r)=j_l(kr)`$
Forma asimtoticฤ este (folosind ec. 51)
$`R_l(r)\frac{1}{2ikr}[e^{ikrl\pi /2}e^{ikrl\pi /2}].`$
Problema 7.3
Aproximaลฃia Born pentru potenลฃiale Yukawa
Sฤ considerฤm un potenลฃial de forma:
$$V(๐ซ)=V_0\frac{e^{\alpha r}}{r},$$
(53)
cu $`V_0`$ ลi $`\alpha `$ constante reale ลi $`\alpha `$ pozitivฤ . Potenลฃialul este atractiv sau repulsiv รฎn funcลฃie de semnul lui $`V_0`$; cu cรฎt este mai mare $`|V_0|`$, cu atรฎt este mai intens potenลฃialul. Presupunem cฤ $`|V_0|`$ este suficient de mic pentru ca aproximaลฃia Born sฤ funcลฃioneze. Conform formulei obลฃinute anterior, amplitudinea de dispersie este datฤ de:
$`f^{(B)}(\theta ,\phi )=\frac{1}{4\pi }\frac{2mV_0}{\mathrm{}^2}e^{i๐๐ซ}\frac{e^{\alpha r}}{r}d^3r.`$
Cum acest potenลฃial depinde numai de $`r`$, integrฤrile unghiulare se pot face uลor, ajungรฎnd astfel la forma:
$`f^{(B)}(\theta ,\phi )=\frac{1}{4\pi }\frac{2mV_0}{\mathrm{}^2}\frac{4\pi }{|๐|}_0^{\mathrm{}}\mathrm{sin}|๐|r\frac{e^{\alpha r}}{r}rdr.`$
Aลadar, obลฃinem:
$`f^{(B)}(\theta ,\phi )=\frac{2mV_0}{\mathrm{}^2}\frac{1}{\alpha ^2+|๐|^2}.`$
Din figurฤ se observฤ cฤ : $`|๐|=2k\mathrm{sin}\frac{\theta }{2}`$; prin urmare:
$`\sigma ^{(B)}(\theta )=\frac{4m^2V_0^2}{\mathrm{}^4}\frac{1}{[\alpha ^2+4k^2\mathrm{sin}\frac{\theta }{2}^2]^2}.`$
Secลฃiunea totalฤ se obลฃine prin integrare:
$`\sigma ^{(B)}=\sigma ^{(B)}(\theta )๐\mathrm{\Omega }=\frac{4m^2V_0^2}{\mathrm{}^4}\frac{4\pi }{\alpha ^2(\alpha ^2+4k^2)}.`$
8. UNDE PARลขIALE
## Introducere
Metoda undelor parลฃiale se referฤ la particule care interacลฃioneazฤ รฎntr-o regiune restrรฎnsฤ de spaลฃiu cu o alta, care prin caracteristicile sale este cunoscutฤ ca centru de รฎmprฤลtiere (de exemplu faptul cฤ se poate considera fixฤ ). รn afara acestei regiuni, interacลฃia รฎntre cele douฤ particule se poate considera neglijabilฤ . รn acest fel este posibil sฤ se descrie particula รฎmprฤลtiatฤ cu urmฤtorul Hamiltonian:
$$H=H_0+V,$$
(1)
unde $`H_0`$ corespunde Hamiltonianului de particulฤ liberฤ . Deci problema noastrฤ este de a rezolva urmฤtoarea ecuaลฃie:
$$(H_0+V)\psi =E\psi .$$
(2)
Este evident cฤ spectrul va fi continuu (studiem cazul รฎmprฤลtierii elastice). Soluลฃia ecuaลฃiei precedente este datฤ de:
$$\psi =\frac{1}{EH_0}V\psi +\varphi .$$
(3)
Cu o analizฤ uลoarฤ putem sฤ vedem cฤ pentru $`V=0`$ obลฃinem soluลฃia $`\varphi `$, adicฤ , soluลฃia corespunzฤtoare particulei libere. Trebuie notat cฤ operatorul $`\frac{1}{EH_0}`$ รฎntr-un anumit sens este anomal, pentru cฤ are un continuu de poli pe axa realฤ care coincid cu valorile proprii ale lui $`H_0`$. Pentru a โscฤpaโ de aceastฤ problemฤ sฤ producem o micฤ deplasare รฎn direcลฃia imaginarฤ ($`\pm iฯต`$) a tฤieturii de pe axa realฤ :
$$\psi ^\pm =\frac{1}{EH_0\pm i\epsilon }V\psi ^\pm +\varphi $$
(4)
Aceastฤ ecuaลฃie este cunoscutฤ ca ecuaลฃia Lippmann-Schwinger. รn final deplasarea polilor va fi รฎn sens pozitiv de la axa imaginarฤ (pentru ca principiul de cauzalitate sฤ nu fie violat \[cf. Feynman\]). Sฤ luฤm reprezentarea x:
$$๐ฑ\psi ^\pm =๐ฑ\varphi +d^3x^{^{}}๐ฑ|\frac{1}{EH_0\pm i\epsilon }|๐ฑ^{^{}}๐ฑ^{^{}}V\psi ^\pm .$$
(5)
Primul termen din partea dreaptฤ corespunde unei particule libere รฎn timp ce al doilea termen se interpreteazฤ ca o undฤ sfericฤ care โieseโ din centrul de รฎmprฤลtiere. Nucleul integralei anterioare se poate asocia cu o funcลฃie Green (sau propagator) ลi este foarte simplu sฤ se calculeze:
$$G_\pm (๐ฑ,๐ฑ^{^{}})=\frac{\mathrm{}^2}{2m}๐ฑ|\frac{1}{EH_0\pm i\epsilon }|๐ฑ^{^{}}=\frac{1}{4\pi }\frac{e^{\pm ik๐ฑ๐ฑ^{^{}}}}{๐ฑ๐ฑ^{^{}}},$$
(6)
unde $`E=\mathrm{}^2k^2/2m`$. Aลa cum am vฤzut mai รฎnainte funcลฃia de undฤ se poate scrie ca o undฤ planฤ plus una sfericฤ care iese din centrul de รฎmprฤลtiere (pรฎnฤ la un factor constant):
$$๐ฑ\psi ^+=e^{๐ค๐ฑ}+\frac{e^{ikr}}{r}f(๐ค,๐ค^{^{}}).$$
(7)
Mฤrimea $`f(๐ค,๐ค^{^{}})`$ care apare รฎn ec. 7 se cunoaลte ca amplitudine de dispersie ลi se poate scrie explicit รฎn forma:
$$f(๐ค,๐ค^{^{}})=\frac{1}{4\pi }(2\pi )^3\frac{2m}{\mathrm{}^2}๐ค^{^{}}V\psi ^+.$$
(8)
Sฤ definim acum un operator T astfel cฤ :
$$T\varphi =V\psi ^+$$
(9)
Dacฤ multiplicฤm ecuaลฃia Lippmann-Schwinger cu V ลi folosim definiลฃia anterioarฤ obลฃinem:
$$T\varphi =V\varphi +V\frac{1}{EH_0+i\epsilon }T\varphi .$$
(10)
Iterรฎnd ecuaลฃia anterioarฤ (ca รฎn teoria de perturbaลฃii) putem obลฃine aproximaลฃia Born ลi corecลฃiile sale de ordin superior.
## Metoda undelor parลฃiale
Sฤ considerฤm acum cazul unui potenลฃial central nenul. รn acest caz, pe baza definiลฃiei (9) se deduce cฤ operatorul $`T`$ comutฤ cu $`\stackrel{}{L}^2`$ ลi $`\stackrel{}{L}`$; de aici se spune cฤ $`T`$ este un operator scalar. รn acest fel pentru a uลura calculele este convenabil sฤ se foloseascฤ coordonate sferice, pentru cฤ datฤ simetria problemei, operatorul $`T`$ va fi diagonal. Acum, sฤ vedem ce formฤ ia expresia (8) pentru amplitudinea de dispersie:
$$f(๐ค,๐ค^{^{}})=\mathrm{const}.\underset{lml^{^{}}m^{^{}}}{}๐E๐E^{^{}}๐ค^{^{}}E^{^{}}l^{^{}}m^{^{}}E^{^{}}l^{^{}}m^{^{}}TElmElm๐ค,$$
(11)
unde $`\mathrm{const}.=\frac{1}{4\pi }\frac{2m}{\mathrm{}^2}(2\pi )^3`$. Dupฤ cรฎteva calcule se obลฃine:
$$f(๐ค,๐ค^{^{}})=\frac{4\pi ^2}{k}\underset{l}{}\underset{m}{}T_l(E)Y_l^m(๐ค^{^{}})Y_l^m^{}(๐ค).$$
(12)
Alegรฎnd sistemul de coordonate astfel ca vectorul $`๐ค`$ sฤ aibฤ aceeaลi direcลฃie cu axa orientatฤ z, se ajunge la concluzia cฤ la amplitudinea de dispersie vor contribui numai armonicele sferice cu m egal cu zero; dacฤ definim $`\theta `$ ca unghiul รฎntre $`๐ค`$ ลi $`๐ค^{^{}}`$ vom avea:
$$Y_l^0(๐ค^{^{}})=\sqrt{\frac{2l+1}{4\pi }}P_l(cos\theta ).$$
(13)
Cu urmฤtoarea definiลฃie:
$$f_l(k)\frac{\pi T_l(E)}{k},$$
(14)
ec. (12) se poate scrie รฎn forma urmฤtoare:
$$f(๐ค,๐ค^{^{}})=f(\theta )=\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)f_l(k)P_l(cos\theta ).$$
(15)
Pentru $`f_l(k)`$ se poate da o interpretare simplฤ pe baza dezvoltฤrii unei unde plane รฎn unde sferice. Astfel putem scrie funcลฃia $`๐ฑ\psi ^+`$ pentru valori mari ale lui $`r`$ รฎn forma
$$๐ฑ\psi ^+=\frac{1}{(2\pi )^{3/2}}\left[e^{ikz}+f(\theta )\frac{e^{ikr}}{r}\right]=$$
$$\frac{1}{(2\pi )^{3/2}}\left[\underset{l}{}(2l+1)P_l(\mathrm{cos}\theta )\left(\frac{e^{ikr}e^{i(krl\pi )}}{2ikr}\right)+\underset{l}{}(2l+1)f_l(k)P_l(\mathrm{cos}\theta )\frac{e^{ikr}}{r}\right]$$
$$=\frac{1}{(2\pi )^{3/2}}\underset{l}{}(2l+1)\frac{P_l(\mathrm{cos}\theta )}{2ik}\left[\left[1+2ikf_l(k)\right]\frac{e^{ikr}}{r}\frac{e^{i(krl\pi )}}{r}\right].$$
(16)
Aceastฤ expresie se poate interpreta dupฤ cum urmeazฤ . Cei doi termeni exponenลฃiali corespund unor unde sferice, primul unei unde emergente ลi al doilea uneia convergente; รฎn plus efectul de รฎmprฤลtiere se vede convenabil รฎn coeficientul undei emergente, care este egal cu unu cรฎnd nu existฤ dispersor.
## Deplasฤri (ลifturi) de fazฤ
Sฤ ne imaginฤm acum o suprafaลฃฤ รฎnchisฤ centratฤ รฎn dispersor. Dacฤ presupunem cฤ nu existฤ creare ลi nici anihilare de particule se verificฤ :
$$๐ฃ๐๐=0,$$
(17)
unde regiunea de integrare este suprafaลฃ definitฤ mai รฎnainte ลi $`๐ฃ`$ este densitatea de curent de probabilitate. รn plus, datoritฤ conservฤrii momentului cinetic ecuaลฃia anterioarฤ trebuie sฤ se verifice pentru fiecare undฤ parลฃialฤ (cu alte cuvinte, toate undele parลฃiale au diferite valori ale proiecลฃiilor momentului cinetic, ceea ce le face diferite. Formularea teoreticฤ ar fi echivalentฤ dacฤ se considerฤ pachetul de unde ca un flux de particule care nu interacลฃioneazฤ รฎntre ele; mai mult, pentru cฤ potenลฃialul problemei noastre este central, momentul cinetic al fiecฤrei โparticuleโ se va conserva ceea ce ne permite sฤ spunem cฤ particulele continuฤ sฤ fie aceleaลi). Cu aceste consideraลฃii, putem sฤ afirmฤm cฤ atรฎt unda divergentฤ cรฎt ลi cea emergentฤ diferฤ cel mult printr-un factor de fazฤ . Deci, dacฤ definim:
$$S_l(k)1+2ikf_l(k)$$
(18)
trebuie sฤ avem
$$S_l(k)=1.$$
(19)
Rezultatele anterioare se pot interpreta cu ajutorul conservฤrii probabilitฤลฃilor ลi erau de โaลteptatโ pentru cฤ nu am presupus cฤ existฤ creare ลi anihilare de particule, astfel cฤ influenลฃa centrului dispersor consistฤ pur ลi simplu รฎn a adฤuga un factor de fazฤ รฎn componentele undelor emergente ลi รฎn virtutea unitaritฤลฃii factorului de fazฤ รฎl putem scrie รฎn forma:
$$S_l=e^{2i\delta _l},$$
(20)
unde $`\delta _l`$ este real ลi este funcลฃie de k. Pe baza definiลฃiei (18) putem sฤ scriem:
$$f_l=\frac{e^{2i\delta _l}1}{2ik}=\frac{e^{i\delta _l}\mathrm{sin}(\delta _l)}{k}=\frac{1}{k\mathrm{cot}(\delta _l)ik}.$$
(21)
Secลฃiunea total de รฎmprฤลtiere ia forma urmฤtoare:
$$\sigma _{total}=f(\theta )^2๐\mathrm{\Omega }=$$
$$\frac{1}{k^2}_0^{2\pi }๐\varphi _1^1d(\mathrm{cos}(\theta ))\underset{l}{}\underset{l^{^{}}}{}(2l+1)(2l^{^{}}+1)e^{i\delta _l}\mathrm{sin}(\delta _l)e^{i\delta _l^{^{}}}\mathrm{sin}(\delta _l^{^{}})P_lP_l^{^{}}$$
$$=\frac{4\pi }{k^2}\underset{l}{}(2l+1)\mathrm{sin}{}_{}{}^{2}(\delta _l^{^{}}).$$
(22)
## Determinarea ลifturilor de fazฤ
Sฤ considerฤm acum un potenลฃial V astfel cฤ se anulฤ pentru $`r>R`$, unde parametrul R se cunoaลte ca โrazฤ de acลฃiune a potenลฃialuluiโ, astfel cฤ regiunea $`r>R`$ evident trebuie sฤ corespundฤ unei unde sferice liberฤ (neperturbatฤ ). Pe de altฤ parte, forma cea mai generalฤ de dezvoltare a unei unde plane รฎn unde sferice este:
$$๐ฑ\psi ^+=\frac{1}{(2\pi )^{3/2}}\underset{l}{}i^l(2l+1)A_l(r)P_l(\mathrm{cos}\theta )(r>R),$$
(23)
unde coeficientul $`A_l`$ este prin definiลฃie:
$$A_l=c_l^{(1)}h_l^{(1)}(kr)+c_l^{(2)}h_l^{(2)}(kr),$$
(24)
ลi unde $`h_l^{(1)}`$ ลi $`h_l^{(2)}`$ sunt funcลฃiile Hankel sferice ale cฤror forme asimptotice sunt:
$$h_l^{(1)}\frac{e^{i(krl\pi /2)}}{ikr}$$
$$h_l^{(2)}\frac{e^{i(krl\pi /2)}}{ikr}.$$
Examinรฎnd forma asimptoticฤ a expresiei (23) care este:
$$\frac{1}{(2\pi )^{3/2}}\underset{l}{}(2l+1)P_l\left[\frac{e^{ikr}}{2ikr}\frac{e^{i(krl\pi )}}{2ikr}\right],$$
(25)
se poate vedea cฤ :
$$c_l^{(1)}=\frac{1}{2}e^{2i\delta _l}c_l^{(2)}=\frac{1}{2}.$$
(26)
Aceasta permite scrierea funcลฃiei de undฤ radialฤ pentru $`r>R`$ รฎn forma:
$$A_l=e^{2i\delta _l}\left[\mathrm{cos}\delta _lj_l(kr)\mathrm{sin}\delta _ln_l(kr)\right].$$
(27)
Folosind ecuaลฃia anterioarฤ putem evalua evalua derivata sa logaritmicฤ รฎn r=R, i.e., exact la frontiera zonei de acลฃiune a potenลฃialului:
$$\beta _l\left(\frac{r}{A_l}\frac{dA_l}{dr}\right)_{r=R}=kR\left[\frac{j_l^{^{}}\mathrm{cos}\delta _ln_l^{^{}}(kR)\mathrm{sin}\delta _l}{j_l\mathrm{cos}\delta _ln_l(kR)\mathrm{sin}\delta _l}\right].$$
(28)
$`j_l^{^{}}`$ este derivata lui $`j_l`$ รฎn raport cu $`kr`$ ลi evaluatฤ รฎn $`r=R`$. Alt rezultat important pe care รฎl putem obลฃine cunoscรฎnd resultatul anterior este ลiftul de fazฤ :
$$\mathrm{tan}\delta _l=\frac{kRj_l^{^{}}(kR)\beta _lj_l(kR)}{kRn_l^{^{}}(kR)\beta _ln_l(kR)}.$$
(29)
Pentru a obลฃine soluลฃia completa a problemei รฎn acest caz este necesar sฤ se facฤ calculele pentru $`r<R`$, adicฤ , รฎn interiorul razei de acลฃiune al potenลฃialului. Pentru cazul unui potenลฃial central, ecuaลฃia Schrรถdinger รฎn trei dimensiuni este:
$$\frac{d^2u_l}{dr^2}+\left(k^2\frac{2m}{\mathrm{}^2}V\frac{l(l+1)}{r^2}\right)u_l=0,$$
(30)
unde $`u_l=rA_l(r)`$ este supusฤ condiลฃiei de frontierฤ $`u_l_{r=0}=0`$. Astfel, putem calcula derivata logaritmicฤ, care รฎn virtutea proprietฤลฃii de continuitate a derivatei logaritmice (care este echivalentฤ cu continuitatea derivatei รฎntr-un punct de discontinuitate) ne conduce la:
$$\beta _l_{interior}=\beta _l_{exterior}.$$
(31)
## Un exemplu: รฎmprฤลtierea pe o sferฤ solidฤ
Sฤ tratฤm acum un caz specific. Fie un potenลฃial definit prin:
$$V=\{\begin{array}{cc}\mathrm{}\hfill & \text{ }r<R\hfill \\ 0\hfill & r>R.\hfill \end{array}$$
(32)
Se ลtie cฤ o particulฤ nu poate penetra รฎntr-o regiune unde potenลฃialul este infinit, astfel cฤ funcลฃia de undฤ trebuie sฤ se anuleze รฎn $`r=R`$; din faptul cฤ sfera este impenetrabilฤ rezultฤ deasemenea cฤ :
$$A_l(r)_{r=R}=0.$$
(33)
Astfel, din ec. (27) avem:
$$\mathrm{tan}\delta _l=\frac{j_l(kR)}{n_l(kR)}.$$
(34)
Se vede cฤ se poate calcula uลor ลiftul de fazฤ pentru orice $`l`$. Sฤ considerฤm acum cazul $`l=0`$ (รฎmprฤลtiere de undฤ s) pentru care avem:
$$\delta _l=kR$$
ลi din ec. (27):
$$A_{l=0}(r)\frac{\mathrm{sin}kr}{kr}\mathrm{cos}\delta _0+\frac{\mathrm{cos}kr}{kr}\mathrm{sin}\delta _0=\frac{1}{kr}\mathrm{sin}(kr+\delta _0).$$
(35)
Vedem cฤ faลฃฤ de miลcarea liberฤ existฤ o contribuลฃie adiลฃionalฤ de o fazฤ . Este clar cฤ รฎntr-un caz mai general diferitele unde vor avea diferite ลifturi de fazฤ ceea ce provoacฤ o distorsie tranzitorie รฎn pachetul de unde dispersat. Sฤ studiem acum cazul energiilor mici, i.e., $`kR<<1`$. รn acest caz, expresiile pentru funcลฃiile Bessel (folosite pentru a descrie funcลฃiile Hankel sferice) sunt urmฤtoarele:
$$j_l(kr)\frac{(kr)^l}{(2l+1)!!}$$
(36)
$$n_l(kr)\frac{(2l1)!!}{(kr)^{l+1}}$$
(37)
care ne conduc la:
$$\mathrm{tan}\delta _l=\frac{(kR)^{2l+1}}{(2l+1)[(2l1)!!]^2}.$$
(38)
Din aceastฤ formulฤ putem sฤ vedem cฤ o contribuลฃie apreciabilฤ la ลiftul de fazฤ este dat de undele cu $`l=0`$ ลi cum $`\delta _0=kR`$ obลฃinem pentru secลฃiunea eficace:
$$\sigma _{total}=\frac{d\sigma }{d\mathrm{\Omega }}๐\mathrm{\Omega }=4\pi R^2.$$
(39)
De aici se ajunge la concluzia cฤ secลฃiunea eficace de รฎmprฤลiere cuanticฤ este de patru ori mai mare decรฎt secลฃiunea eficace clasicฤ ลi coincide cu aria totalฤ a sferei dure. Pentru valori mari ale energiei pachetului incident se poate lucra cu ipoteza cฤ toate valorile lui $`l`$ pรฎnฤ la o valoare maximฤ $`l_{max}kR`$ contribuie la secลฃiunea eficace totalฤ :
$$\sigma _{total}=\frac{4\pi }{k^2}\underset{l=0}{\overset{lkR}{}}(2l+1)\mathrm{sin}^2\delta _l.$$
(40)
รn acest fel, pe baza ec. (34) avem:
$$\mathrm{sin}^2\delta _l=\frac{\mathrm{tan}^2\delta _l}{1+\mathrm{tan}^2\delta _l}=\frac{[j_l(kR)]^2}{[j_l(kR)]^2+[n_l(kR)]^2}\mathrm{sin}^2\left(kR\frac{l\pi }{2}\right),$$
(41)
unde am folosit expresiile:
$$j_l(kr)\frac{1}{kr}\mathrm{sin}\left(kr\frac{l\pi }{2}\right)$$
$$n_l(kr)\frac{1}{kr}\mathrm{cos}\left(kr\frac{l\pi }{2}\right).$$
Vedem cฤ $`\delta _l`$ descreลte cu $`\frac{\pi }{2}`$ de fiecare datฤ cฤ $`l`$ se incrementฤ cu o unitate, ลi deci este evident cฤ se รฎndeplineลte $`\mathrm{sin}^2\delta _l+\mathrm{sin}^2\delta _{l+1}=1`$. Aproximรฎnd $`\mathrm{sin}^2\delta _l`$ cu valoarea sa medie $`\frac{1}{2}`$, este simplu de obลฃinut rezultatul pe baza sumei de numere impare:
$$\sigma _{total}=\frac{4\pi }{k^2}(kR)^2\frac{1}{2}=2\pi R^2.$$
(42)
Odatฤ รฎn plus rezultatul calculului bazat pe metodele de mecanicฤ cuanticฤ , deลi asemฤnฤtor, diferฤ totuลi de rezultatul clasic. Sฤ vedem care este originea factorului 2; mai รฎntรฎi vom separa ec. (15) รฎn douฤ pฤrลฃi:
$$f(\theta )=\frac{1}{2ik}\underset{l=0}{\overset{l=kR}{}}(2l+1)e^{2i\delta _l}P_l\mathrm{cos}(\theta )+\frac{i}{2k}\underset{l=0}{\overset{l=kR}{}}(2l+1)P_l\mathrm{cos}(\theta )=f_{\text{refl}}+f_{\text{umbrฤ }}.$$
(43)
Evaluรฎnd $`|f_{\text{ refl}}|^2๐\mathrm{\Omega }`$:
$$|f_{\text{ refl}}|^2๐\mathrm{\Omega }=\frac{2\pi }{4k^2}\underset{l=0}{\overset{l_{max}}{}}_{1}^{}{}_{}{}^{1}(2l+1)^2[P_l\mathrm{cos}(\theta )]^2d(\mathrm{cos}\theta )=\frac{\pi l_{max}^{}{}_{}{}^{2}}{k^2}=\pi R^2.$$
(44)
Analizรฎnd acum $`f_{\text{umbrฤ }}`$ pentru unghiuri mici avem:
$$f_{\text{umbrฤ }}\frac{i}{2k}(2l+1)J_0(l\theta )ik_0^RbJ_0(kb\theta )๐b=\frac{iRJ_1(kR\theta )}{\theta }.$$
(45)
Aceastฤ formulฤ este destul de cunoscutฤ รฎn opticฤ , fiind formula pentru difracลฃia Fraunhofer; cu ajutorul schimbului de variabilฤ $`z=kR\theta `$ putem sฤ evaluฤm integrala $`|f_{\text{ umbrฤ }}|^2๐\mathrm{\Omega }`$:
$$|f_{\text{ umbrฤ }}|^2๐\mathrm{\Omega }2\pi R^2_0^{\mathrm{}}\frac{[J_1(z)]^2}{z}๐z\pi R^2.$$
(46)
รn sfรฎrลit, neglijรฎnd interferenลฃa รฎntre $`f_{\text{refl}}`$ ลi $`f_{\text{ umbrฤ }}`$ (pentru cฤ faza oscileazฤ รฎntre $`2\delta _{l+1}=2\delta _l\pi `$). Se obลฃine astfel rezultatul (42). Am etichetat unul dintre termeni cu titlul de umbrฤ , pentru cฤ originea sa se explicฤ uลor dacฤ se apeleazฤ la comportamentul ondulatoriu al particulei dispersate (din punct de vedere โfizicโ nu existฤ nici o diferenลฃฤ รฎntre un pachet de undฤ ลi o particulฤ รฎn acest caz). Originea sa constฤ รฎn componentele pachetului de unde รฎmprฤลtiate รฎnapoi ceea ce produce o diferenลฃฤ de fazฤ faลฃฤ de undele incidente ducรฎnd la o interferenลฃฤ distructivฤ .
## รmprฤลtiere รฎn cรฎmp coulombian
Sฤ considerฤm acum un exemplu clasic ลi ceva mai complicat: รฎmprฤลtierea de particule รฎntr-un cรฎmp coulombian. Pentru acest caz ecuaลฃia Schrรถdinger este:
$$\left(\frac{\mathrm{}^2}{2m}^2\frac{Z_1Z_2e^2}{r}\right)\psi (๐ซ)=E\psi (๐ซ),E>0,$$
(47)
unde $`m`$ este masa redusฤ a sistemului รฎn interacลฃie ลi evident $`E>0`$ deoarece tratฤm cazul dispersiei fฤrฤ producere de nici un fel de stฤri legate. Ecuaลฃia anterioarฤ este echivalentฤ urmฤtoarei expresii (pentru valori adecuate ale constantelor $`k`$ ลi $`\gamma `$) :
$$\left(^2+k^2+\frac{2\gamma k}{r}\right)\psi (๐ซ)=0.$$
(48)
Dacฤ nu considerฤm bariera centrifugalฤ a potenลฃialului efectiv (unde $`s`$) ne gฤsim รฎn condiลฃiile unei interacลฃiuni coulombiene pure ลi putem propune o soluลฃie de forma:
$$\psi (๐ซ)=e^{i๐ค๐ซ}\chi (u),$$
(49)
cu
$$u=ikr(1\mathrm{cos}\theta )=ik(rz)=ikw,$$
$$๐ค๐ซ=kz.$$
$`\psi (๐ซ)`$ este soluลฃia completฤ a ecuaลฃiei Schrรถdinger ลi se poate aลtepta un comportament asimptotic format din douฤ pฤrลฃi, respectiv de undฤ planฤ $`e^{i๐ค๐ซ}`$ ลi undฤ sfericฤ $`r^1e^{ikr}`$. Definind noi variabile:
$$z=zw=rz\lambda =\varphi ,$$
cu ajutorul relaลฃiilor anterioare, ec. (48) ia forma:
$$\left[u\frac{d^2}{du^2}+(1u)\frac{d}{du}i\gamma \right]\chi (u)=0.$$
(50)
Pentru a rezolva aceastฤ ecuaลฃie, trebuie studiat mai รฎntรฎi comportamentul sฤu asimptotic, dar cum acesta a fost deja prezentat, funcลฃia de undฤ asimptoticฤ normalizatฤ care se obลฃine รฎn final ca rezultat al tuturor calculelor anterioare este:
$$\psi _๐ค(๐ซ)=\frac{1}{(2\pi )^{3/2}}\left(e^{i[๐ค๐ซ\gamma ln(kr๐ค๐ซ)]}+\frac{f_c(k,\theta )e^{i[kr+\gamma ln2kr]}}{r}\right).$$
(51)
Dupฤ cum vedem, funcลฃia de undฤ anterioarฤ prezintฤ termeni care o fac sฤ difere apreciabil de ec. (7). Acest lucru se datoreazฤ faptului cฤ forลฃa coulombianฤ este de razฤ infinitฤ de acลฃiune. Efectuarea calculului exact pentru amplitudinea de รฎmprฤลฃiere coulombianฤ este destul de dificil de realizat. Aici vom da numai rezultatul final pentru funcลฃia de undฤ normalizatฤ :
$$\psi _๐ค(๐ซ)=\frac{1}{(2\pi )^{3/2}}\left(e^{i[๐ค๐ซ\gamma ln(kr๐ค๐ซ)]}+\frac{g_1^{}(\gamma )}{g_1(\gamma )}\frac{\gamma }{2k\mathrm{sin}(\theta /2)^2}\frac{e^{i[kr+\gamma ln2kr]}}{r}\right),$$
(52)
unde $`g_1(\gamma )=\frac{1}{\mathrm{\Gamma }(1i\gamma )}`$.
รn ceea ce priveลte analiza de unde parลฃiale o vom reduce la prezentarea rezultatelor deja discutate รฎntr-un mod cรฎt mai clar posibil. Mai รฎntรฎi scriem funcลฃia de undฤ (49) $`\psi (๐ซ)`$ รฎn urmฤtoarea formฤ :
$$\psi (๐ซ)=e^{i๐ค๐ซ}\chi (u)=Ae^{i๐ค๐ซ}_Ce^{ut}t^{i\gamma 1}(1t)^{i\gamma }๐t,$$
(53)
unde $`A`$ este o constantฤ de normalizare ลi toatฤ partea integralฤ este transformata Laplace inversฤ a transformatei directe a ecuaลฃiei (50). O formฤ convenabilฤ a ecuaลฃiei anterioare este:
$$\psi (๐ซ)=A_Ce^{i๐ค๐ซ}(1t)e^{ikrt}(1t)d(t,\gamma )๐t$$
(54)
cu
$$d(t,\gamma )=t^{i\gamma 1}(1t)^{i\gamma 1}.$$
(55)
รn cadrul analizei de unde parลฃiale procedฤm la a scrie:
$$\psi (๐ซ)=\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)i^lP_l(\mathrm{cos}\theta )A_l(kr),$$
(56)
unde
$$A_l(kr)=A_Ce^{ikrt}j_l[kr(1t)](1t)d(t,\gamma ).$$
(57)
Aplicรฎnd relaลฃiile รฎntre funcลฃiile Bessel sferice ลi funcลฃiile Hankel sferice avem:
$$A_l(kr)=A_l^{(1)}(kr)+A_l^{(2)}(kr).$$
(58)
Evaluarea acestor coeficienลฃi nu o vom prezenta aici (fiind destul de complicatฤ ). Rezultฤ cฤ :
$$A_l^{(1)}(kr)=0$$
(59)
$$A_l^{(2)}(kr)\frac{Ae^{\pi \gamma /2}}{2ikr}[2\pi ig_1(\gamma )]\left(e^{i[kr(l\pi /2)+\gamma \mathrm{ln}2kr]}e^{2i\eta _l(k)}e^{i[kr(l\pi /2)+\gamma \mathrm{ln}2kr]}\right)$$
(60)
unde
$$e^{2i\eta _l(k)}=\frac{\mathrm{\Gamma }(1+li\gamma )}{\mathrm{\Gamma }(1+l+i\gamma )}.$$
(61)
## Calculul amplitudinii de รฎmprฤลtiere coulombianฤ
Dacฤ efectuฤm transformata Laplace a ec. (50) obลฃinem:
$$\chi (u)=A_Ce^{ut}t^{i\gamma 1}(1t)^{i\gamma }๐t.$$
(62)
Conturul $`C`$ merge de la $`\mathrm{}`$ la $`\mathrm{}`$ ลi se รฎnchide pe deasupra axei reale. รn aceste condiลฃii, vedem cฤ existฤ doi poli: cรฎnd $`t=0`$ ลi $`t=1`$. Cu schimbul de variabilฤ $`s=ut`$ obลฃinem:
$$\chi (u)=A_{C_1}e^ss^{i\gamma 1}(us)^{i\gamma }.$$
(63)
$`\chi (u)`$ trebuie sฤ fie regularฤ รฎn zero ลi รฎntr-adevฤr:
$$\chi (0)=(1)^{i\gamma }A_{C_1}\frac{e^s}{s}ds.=(1)^{i\gamma }A2\pi i$$
(64)
Luรฎnd acum limita $`u\mathrm{}`$, sฤ facem o deplasare infinitezimalฤ (pentru a elimina faptul cฤ polii sunt pe contur) ลi cu un schimb de variabilฤ astfel cฤ $`\frac{s}{u}=\frac{(s_0\pm i\epsilon )}{i\kappa }`$, vedem cฤ aceastฤ expresie tinde la zero cรฎnd $`u\mathrm{}`$. Deci, putem sฤ dezvoltฤm $`(us)`$ รฎn serie de puteri de $`\frac{s}{u}`$ pentru polul cu $`s=0`$. Dar aceastฤ dezvoltare nu este bunฤ รฎn $`s=1`$, pentru cฤ รฎn acest caz $`s=s_0+i(\kappa \pm \epsilon )`$ ลi de aici rezultฤ cฤ $`\frac{s}{u}=1\frac{(s_0\pm i\epsilon )}{\kappa }`$ tinde la $`1`$ cรฎnd $`\kappa \mathrm{}`$; dar dacฤ facem schimbul de variabilฤ $`s^{^{}}=su`$ aceastฤ problemฤ se eliminฤ :
$$\chi (u)=A_{\mathrm{C}_2}\left([e^ss^{i\gamma 1}(us)^{i\gamma }]ds+[e^{s^{^{}}+u}(s^{^{}})^{i\gamma }(u+s^{^{}})^{i\gamma 1}]ds^{^{}}\right).$$
(65)
Dezvoltรฎnd seriile de puteri este uลor de calculat integralele precedente, dar รฎn rezultat trebuie sฤ se ia limita $`\frac{s}{u}0`$ pentru a obลฃine formele asimptotice corecte pentru รฎmprฤลtierea coulombianฤ :
$$\chi (u)2\pi iA\left[u^{i\gamma }g_1(\gamma )(u)^{i\gamma 1}e^ug_2(\gamma )\right]$$
$$2\pi g_1(\gamma )=i_{\mathrm{C}_2}e^ss^{i\gamma 1}๐s$$
$$2\pi g_2(\gamma )=i_{\mathrm{C}_2}e^ss^{i\gamma }๐s.$$
(66)
Dupฤ acest ลir de schimbฤri de variabile, ne รฎntoarcem la $`s`$-ul original pentru a obลฃine:
$$(u^{})^{i\gamma }=(i)^{i\gamma }[k(rz)]^{i\gamma }=e^{\gamma \pi /2}e^{i\gamma \mathrm{ln}k(rz)}$$
$$(u)^{i\gamma }=(i)^{i\gamma }[k(rz)]^{i\gamma }=e^{\gamma \pi /2}e^{i\gamma \mathrm{ln}k(rz)}.$$
(67)
Calculul lui $`\chi `$ odatฤ efectuat, este echivalent cu a avea $`\psi _๐ค(๐ซ)`$ pornind de la (49).
## Aproximaลฃia eikonalฤ
Vom face o scurtฤ expoziลฃie a aproximaลฃiei eikonale a cฤrei filosofie este aceeaลi cu cea care se face cรฎnd se trece de la optica ondulatorie la optica geometricฤ ลi de aceea este corectฤ cรฎnd potenลฃialul variazฤ puลฃin pe distanลฃe comparabile cu lungimea de undฤ a pachetului de unde dispersat, adicฤ , pentru cazul $`E>>|V|`$. Astfel aceastฤ aproximaลฃie poate fi consideratฤ ca o aproximaลฃie cuasiclasicฤ . Mai รฎntรฎi propunem cฤ funcลฃia de undฤ cuasiclasicฤ are forma binecunoscutฤ :
$$\psi e^{iS(๐ซ)/\mathrm{}},$$
(68)
unde S satisface ecuaลฃia Hamilton-Jacobi, cu soluลฃia:
$$\frac{S}{\mathrm{}}=_{\mathrm{}}^z\left[k^2\frac{2m}{\mathrm{}^2}V\left(\sqrt{b^2+z^2}\right)\right]^{1/2}๐z^{}+\text{ constantฤ}.$$
(69)
Constanta aditivฤ se alege รฎn aลa fel รฎncรฎt:
$$\frac{S}{\mathrm{}}kz\mathrm{pentru}V0.$$
(70)
Termenul care multiplicฤ potenลฃialul se poate interpreta ca un schimb de fazฤ al pachetului de unde, avรฎnd urmฤtoarea formฤ explicitฤ
$$\mathrm{\Delta }(b)\frac{m}{2k\mathrm{}^2}_{\mathrm{}}^{\mathrm{}}V\left(\sqrt{b^2+z^2}\right)๐z.$$
(71)
รn cadrul metodei de unde parลฃiale aceastฤ aproximaลฃie are urmฤtoarea aplicaลฃie. ลtim cฤ aproximaลฃia eikonalฤ este corectฤ la energii รฎnalte, unde existฤ multe unde parลฃiale care contribuie la dispersie. Astfel putem considera $`l`$ ca o variabilฤ continuฤ ลi prin analogie cu mecanica clasicฤ punem $`l=bk`$. รn plus, cum deja am menลฃionat $`l_{max}=kR`$, care substituit รฎn expresia (15) conduce la:
$$f(\theta )=ikbJ_0(kb\theta )[e^{2i\mathrm{\Delta }(b)}1]๐b.$$
(72)
## 8P. Probleme
Problema 8.1
Sฤ se obลฃinฤ deplasarea de fazฤ (ลiftul) ลi secลฃiunea diferenลฃialฤ de รฎmprฤลtiere la unghiuri mici pentru un centru de รฎmprฤลtiere de potenลฃial $`U(r)=\frac{\alpha }{r^2}`$. Sฤ se ลฃinฤ cont de faptul cฤ รฎn รฎmprฤลtierile de unghiuri mici principala contribuลฃie o dau undele parลฃiale cu $`l`$ mari.
Soluลฃie:
Rezolvรฎnd ecuaลฃia
$$R_l^{^{\prime \prime }}+\left[k^2\frac{l(l+1)}{r^2}\frac{2m\alpha }{\mathrm{}^2r^2}\right]=0$$
cu condiลฃiile de frontierฤ $`R_l(0)=0`$, $`R_l(\mathrm{})=N`$, unde $`N`$ este un numฤr finit, obลฃinem
$$R_l(r)=A\sqrt{r}I_\lambda (kr),$$
unde $`\lambda =\left[(l+\frac{1}{2})^2+\frac{2m\alpha }{\mathrm{}^2}\right]^{1/2}`$ ลi $`I`$ este prima funcลฃia Bessel modificatฤ .
Pentru determinarea lui $`\delta _l`$ se foloseลte expresia asimptoticฤ a lui $`I_\lambda `$:
$$I_\lambda (kr)\left(\frac{2}{\pi kr}\right)^{1/2}\mathrm{sin}(kr\frac{\lambda \pi }{2}+\frac{\pi }{4}).$$
Prin urmare
$$\delta _l=\frac{\pi }{2}\left(\lambda l\frac{1}{2}\right)=\frac{\pi }{2}\left(\left[(l+\frac{1}{2})^2+\frac{2m\alpha }{\mathrm{}^2}\right]^{1/2}\left(l+\frac{1}{2}\right)\right).$$
Condiลฃia $`l`$ mari de care se vorbeลte รฎn problemฤ ne conduce la:
$$\delta _l=\frac{\pi m\alpha }{(2l+1)\mathrm{}^2},$$
de unde se vede cฤ $`|\delta _l|1`$ pentru $`l`$ mari.
Din expresia generalฤ a amplitudinii de รฎmprฤลtiere
$$f(\theta )=\frac{1}{2ik}\underset{l=0}{\overset{\mathrm{}}{}}(2l+1)P_l(\mathrm{cos}\theta )(e^{2i\delta _l}1),$$
la unghiuri mici $`e^{2i\delta _l}1+2i\delta _l`$ ลi
$$\underset{l=0}{\overset{\mathrm{}}{}}P_l(\mathrm{cos}\theta )=\frac{1}{2\mathrm{sin}\frac{\theta }{2}}.$$
Astfel:
$$f(\theta )=\frac{\pi \alpha m}{k\mathrm{}^2}\frac{1}{2\mathrm{sin}\frac{\theta }{2}}.$$
Prin urmare:
$$\frac{d\sigma }{d\theta }=\frac{\pi ^3\alpha ^2m}{2\mathrm{}^2E}\text{c}tg\frac{\theta }{2}.$$ |
warning/0003/gr-qc0003061.html | ar5iv | text | # Behaviour of spin-1/2 particle around a charged black hole
## I. INTRODUCTION
Chandrasekhar separated Dirac equation in Kerr geometry into radial and angular parts in 1976. His separation method can be extended to Schwarzschild geometry and corresponding separated equations can be found. But he did not consider the charge of the black hole. If we consider the black hole as a charged one then electromagnetic interaction is important for incoming particle with charge. To study the behaviour of spin-$`\frac{1}{2}`$ particle, Dirac wave is treated as a perturbation in the space-time which is asymptotically flat . Far away from the black hole its influence on particle is not significant. As it comes closer, feels the curvature of the space-time and corresponding behaviours start to change with respect to that of flat space. Their behaviour around black-hole without charge have been studied in the past by several authors \[1-6\]. In this paper, we will introduce charge in the black hole. Here, we study a simpler problem to have a feeling about the solution when the black hole is non-rotating but charged. Here we have to solve Dirac equation in electromagnetic field around a Reissner-Nordstrรถm black-hole. Thus we will study the particle in crossed electromagnetic and gravitational field. It is very clear that the potential felt by the incoming Dirac wave will be different from that for Schwarzschild black hole . For the incoming uncharged particle like neutron, electromagnetic field does not play any part and the Dirac equation will be reduced to same as Schwarzschild case except the re-definition of horizon. For charged incoming particle like electron, proton etc. electromagnetic gauge field should be introduced. One also can study the neutrino wave whose behaviour is known for Kerr geometry . In the next Section, we present the basic Dirac equations and separate them in this crossed field. In ยง3, we study the behaviour of the potential and possibilities of super-radiance. In ยง4, we present a complete solution. Finally, in ยง5, we draw our conclusions.
## II. DIRAC EQUATION AND ITโS SEPARATION
By introducing electromagnetic interaction and gravitational effect the covariant derivatives take the form as
$$D_\mu =_\mu +iq_1A_\mu +q_2\mathrm{\Gamma }_\mu .$$
$`(1)`$
The derivative of spinor $`P^A`$ can be written as
$$D_\mu P^A=_\mu P^A+iq_1A_\mu P^A+q_2\mathrm{\Gamma }_{\mu \nu }^AP^\nu ,$$
$`(2)`$
where, $`q_1`$ and $`q_2`$ are coupling constants. $`q_1`$ is the charge of the incoming particle (say $`q_1=q`$) and $`q_2`$ is chosen throughout $`1`$. $`A_\mu `$ and $`\mathrm{\Gamma }_{\mu \nu }^A`$ are electromagnetic and gravitational gauge (spin coefficients) fields respectively. Thus, following the Dirac equation in Newman-Penrose formalism can be written as
$$\sigma _{AB^{}}^\mu D_\mu P^A+i\mu _p\overline{Q}^C^{}ฯต_{C^{}B^{}}=0,$$
$`(3a)`$
$$\sigma _{AB^{}}^\mu D_\mu Q^A+i\mu _p\overline{P}^C^{}ฯต_{C^{}B^{}}=0,$$
$`(3b)`$
where, for any vector $`X_i`$, according to the spinor formalism $`\sigma _{AB^{}}^iX_i=X_{AB^{}}`$; $`A,B=0,1`$. Here, we introduce a null tetrad $`(\stackrel{}{l},\stackrel{}{n},\stackrel{}{m},\stackrel{}{\overline{m}})`$ to satisfy orthogonality relations, $`\stackrel{}{l}.\stackrel{}{n}=1`$, $`\stackrel{}{m}.\stackrel{}{\overline{m}}=1`$ and $`\stackrel{}{l}.\stackrel{}{m}=\stackrel{}{n}.\stackrel{}{m}=\stackrel{}{l}.\stackrel{}{\overline{m}}=\stackrel{}{n}.\stackrel{}{\overline{m}}=0`$ following Newman & Penrose . $`2^{\frac{1}{2}}\mu _p`$ is the mass of the Dirac particle. In terms of this new basis in Newman-Penrose formalism Pauli matrices can be written as
$$\sigma _{AB^{}}^\mu =\frac{1}{\sqrt{2}}\left(\begin{array}{cc}l^\mu & \hfill m^\mu \\ \overline{m}^\mu & \hfill n^\mu \end{array}\right).$$
$`(4)`$
Using equation (2), (3a), (4) and choosing $`B=0`$ and subsequently $`B=1`$ we get
$$l^\mu (_\mu +iqA_\mu )P^0+\overline{m}^\mu (_\mu +iqA_\mu )P^1+(\mathrm{\Gamma }_{1000^{}}\mathrm{\Gamma }_{0010^{}})P^0+(\mathrm{\Gamma }_{1100^{}}\mathrm{\Gamma }_{0110^{}})P^1i\mu _p\overline{Q}^1^{}=0,$$
$`(5a)`$
$$m^\mu (_\mu +iqA_\mu )P^0+n^\mu (_\mu +iqA_\mu )P^1+(\mathrm{\Gamma }_{1001^{}}\mathrm{\Gamma }_{0011^{}})P^0+(\mathrm{\Gamma }_{1101^{}}\mathrm{\Gamma }_{0111^{}})P^1+i\mu _p\overline{Q}^0^{}=0,$$
$`(5b)`$
Next by taking complex conjugation of equation (3b), writing various spin coefficients by their named symbol and choosing
$$P^0=F_1,P^1=F_2,\overline{Q}^1^{}=G_1,\overline{Q}^0=G_2$$
we get
$$l^\mu (_\mu +iqA_\mu )F_1+\overline{m}^\mu (_\mu +iqA_\mu )F_2+(ฯต\rho )F_1+(\pi \alpha )F_2=i\mu _pG_1,$$
$`(6a)`$
$$m^\mu (_\mu +iqA_\mu )F_1+n^\mu (_\mu +iqA_\mu )F_2+(\mu \gamma )F_2+(\beta \tau )F_1=i\mu _pG_2,$$
$`(6b)`$
$$l^\mu (_\mu +iqA_\mu )G_2m^\mu (_\mu +iqA_\mu )G_1+(ฯต^{}\rho ^{})G_2(\pi ^{}\alpha ^{})G_1=i\mu _pF_2,$$
$`(6c)`$
$$n^\mu (_\mu +iqA_\mu )G_1\overline{m}^\mu (_\mu +iqA_\mu )G_2+(\mu ^{}\gamma ^{})G_1(\beta ^{}\tau ^{})G_2=i\mu _pF_1,$$
$`(6d)`$
These are the Dirac equations in Newman-Penrose formalism in curved space-time with the presence of electromagnetic interaction.
Now we write the basis vectors of null tetrad in terms of elements of the Reissner-Nordstrรถm geometry as,
$$l^\mu =\frac{1}{\mathrm{\Delta }}(r^2,\mathrm{\Delta },0,0),$$
$`(7a)`$
$$n^\mu =\frac{1}{2r^2}(r^2,\mathrm{\Delta },0,0),$$
$`(7b)`$
$$m^\mu =\frac{1}{r\sqrt{2}}(0,0,1,icosec\theta ),$$
$`(7c)`$
$$\overline{m}^\mu =\frac{1}{r\sqrt{2}}(0,0,1,icosec\theta ),$$
$`(7d)`$
where, $`\mathrm{\Delta }=r^22Mr+Q_{}^2`$ and $`G=\text{}h=c=1`$ are chosen. Here $`M`$ is mass of the black hole, $`Q_{}`$ is charge of the black hole, $`G`$ is gravitational constant, $`h`$ is Plankโs constant, $`c`$ is speed of light.
We consider the spin-$`\frac{1}{2}`$ wave function as the form of $`e^{i(\sigma t+m\varphi )}f(r,\theta )`$ where, $`\sigma `$ is the frequency of the incoming wave and $`m`$ is the azimuthal quantum number. The temporal and azimuthal dependencies are chosen same but radial and polar dependencies are chosen different for different spinors. Thus we write,
$$f_1=e^{i(\sigma t+m\varphi )}rF_1,f_2=e^{i(\sigma t+m\varphi )}F_2,g_1=e^{i(\sigma t+m\varphi )}G_1,g_2=e^{i(\sigma t+m\varphi )}rG_2$$
$`(8)`$
Now we strictly consider the static field so the magnetic potentials are chosen zero, i.e., $`A^\mu =(A^t,0,0,0)`$. $`A^t`$ is nothing but corresponding scaler potential of the field as (in this spherically symmetric space-time)
$$A^t=\frac{qQ_{}}{rr_+},$$
$`(9)`$
where, $`r_+=\mathrm{locationofthehorizon}=M+\sqrt{M^2Q_{}^2}`$.
So using equations (7), (8) and (9) and writing various spin coefficients in terms of the Reissner-Nordstrรถm metric elements (actually in terms of basis vectors) equation (6)s reduce to
$$๐_0f_1+2^{\frac{1}{2}}_{\frac{1}{2}}f_2=i\mu _prg_1$$
$`(10a)`$
$$\mathrm{\Delta }๐_{\frac{1}{2}}^{}f_22^{\frac{1}{2}}_{\frac{1}{2}}^{}f_1=2i\mu _prg_2$$
$`(10b)`$
$$๐_0g_22^{\frac{1}{2}}_{\frac{1}{2}}^{}g_1=i\mu _prf_2$$
$`(10c)`$
$$\mathrm{\Delta }๐_{\frac{1}{2}}^{}g_1+2^{\frac{1}{2}}_{\frac{1}{2}}g_2=2i\mu _prf_1$$
$`(10d)`$
where,
$$๐_n=\frac{d}{dr}+\frac{ir^2\sigma }{\mathrm{\Delta }}+\frac{iqQ_{}r^2}{\mathrm{\Delta }(rr_+)}+2n\frac{rM}{\mathrm{\Delta }},๐_n^{}=\frac{d}{dr}\frac{ir^2\sigma }{\mathrm{\Delta }}\frac{iqQ_{}r^2}{\mathrm{\Delta }(rr_+)}+2n\frac{rM}{\mathrm{\Delta }},$$
$`(11)`$
$$_n=\frac{d}{d\theta }+Q+ncot\theta ,_n^{}=\frac{d}{d\theta }Q+ncot\theta ,Q=mcosec\theta .$$
$`(12)`$
Now considering $`f_1(r,\theta )=R_{\frac{1}{2}}(r)S_{\frac{1}{2}}(\theta )`$, $`f_2(r,\theta )=R_{\frac{1}{2}}(r)S_{\frac{1}{2}}(\theta )`$, $`g_1(r,\theta )=R_{\frac{1}{2}}(r)S_{\frac{1}{2}}(\theta )`$, $`g_2(r,\theta )=R_{\frac{1}{2}}(r)S_{+\frac{1}{2}}(\theta )`$ and following Chandrasekhar we can separate the Dirac equation into radial and angular parts as
$$\mathrm{\Delta }^{1/2}๐_0R_{1/2}=(\lambda +im_pr)\mathrm{\Delta }^{1/2}R_{1/2},$$
$`(13a)`$
$$\mathrm{\Delta }^{1/2}๐_0^{}\mathrm{\Delta }^{1/2}R_{1/2}=(\lambda im_pr)R_{1/2},$$
$`(13b)`$
$$_{1/2}S_{1/2}=\lambda S_{1/2},$$
$`(14a)`$
$$_{1/2}^{}S_{1/2}=\lambda S_{1/2}.$$
$`(14b)`$
Here, $`m_p`$ is the normalised rest mass of the incoming particle and $`\lambda `$ is the sepreration constant.
## III. Nature of the Potential in decoupled system
The equations (14a-b) are same as the angular equation in Schwarzschild geometry whose solution is given in as
$$\lambda ^2=\left(l+\frac{1}{2}\right)^2,R_{\pm \frac{1}{2}}=\mathrm{standaredsphericalharmonics}=_{\pm \frac{1}{2}}Y_m^l(\theta ).$$
$`(15)`$
It is clear that the separation constant depends on orbital angular momentum quantum number $`l`$.
The equations (13a-b) are in coupled form. Following Chandrasekharโs and Mukhopadhyay & Chakrabartiโs approach we can decouple it as
$$\left(\frac{d^2}{d\widehat{r}_{}^2}+\sigma ^2\right)Z_\pm =V_\pm Z_\pm ,$$
$`(16)`$
where,
$$\widehat{r}_{}=r_{}+\frac{1}{2\sigma }tan^1\frac{m_pr}{\lambda }+\frac{qQ_{}}{\sigma }\left[log(rr_{})+\left\{\frac{2r_+}{r_+r_{}}\frac{r_+^2}{(r_+r_{})^2}\right\}log\left(\frac{rr_+}{rr_{}}\right)\frac{r_+^2}{(rr_{})(rr_+)}\right],$$
$`(17)`$
$$r_{}=r3M+\frac{r_+^2}{r_+r_{}}log(rr_+)\frac{r_{}^2}{r_+r_{}}log(rr_{}),$$
$`(18)`$
$$r_\pm =M\pm \sqrt{M^2Q_{}^2},Z_\pm =\mathrm{\Delta }^{1/2}R_{1/2}e^{i\mathrm{\Theta }/2}\pm R_{1/2}e^{i\mathrm{\Theta }/2},\mathrm{\Theta }=m_pr.$$
$`(19)`$
In the extreme case when $`M=Q^{}`$, expression for $`\widehat{r}_{}`$ and $`r_{}`$ are given as,
$$\widehat{r}_{}=r_{}+\frac{1}{2\sigma }tan^1\frac{m_pr}{\lambda }+\frac{qQ_{}}{\sigma }\left[log(rM)\frac{2M^2}{(rM)^2}\frac{2M}{(rM)}\right],$$
$`(17^{})`$
$$r_{}=rM+2Mlog(rM)\frac{M^2}{(rM)}.$$
$`(18^{})`$
Here, $`\widehat{r}_{}`$ is varying from $`\mathrm{}`$ to $`+\mathrm{}`$ (cartesian coordinate). If we compare equation (16) with one dimensional Schrรถdinger equation in cartesian coordinate system the energy $`E`$ of the incoming particle can be written as $`E\sigma ^2`$ and the potential ($`V_\pm `$) felt by the particle is given as
$$V_\pm =\frac{\mathrm{\Delta }(\lambda ^2+m_p^2r^2)^3}{\left[r^2(\lambda ^2+m_p^2r^2)\left(1+\frac{Q_{}q}{(rr_+)\sigma }\right)+\frac{\mathrm{\Delta }\lambda m_p}{2\sigma }\right]^2}\pm \frac{\mathrm{\Delta }(\lambda ^2+m_p^2r^2)}{\left[r^2(\lambda ^2+m_p^2r^2)\left(1+\frac{Q_{}q}{(rr_+)\sigma }\right)+\frac{\mathrm{\Delta }\lambda m_p}{2\sigma }\right]^3}$$
$$[\{r^2(\lambda ^2+m_p^2r^2)(1+\frac{Q_{}q}{(rr_+)\sigma })+\frac{\mathrm{\Delta }\lambda m_p}{2\sigma }\}\frac{(\lambda ^2+m_p^2r^2)^{1/2}}{\mathrm{\Delta }^{1/2}}\{(rM)(\lambda ^2+m_p^2r^2)+3\mathrm{\Delta }rm_p^2\}\mathrm{\Delta }^{1/2}(\lambda ^2+m_p^2r^2)^{3/2}$$
$$\{2r(\lambda ^2+m_p^2r^2)(1+\frac{Q_{}q}{(rr_+)\sigma })+2r^3m_p^2(1+\frac{Q_{}q}{(rr_+)\sigma })r^2(\lambda ^2+m_p^2r^2)\frac{Q_{}q}{(rr_+)^2\sigma }+\frac{(rM)\lambda m_p}{\sigma }\}].$$
$`(20)`$
From the expression of $`V_\pm `$ it is very clear that potential strictly depends on charge of the particle as well as of black hole. More precisely it depends on Coulomb interaction between charge of black hole and incoming particle. When charge of the black hole or particle or both are chosen zero the potential reduces to same as that in Schwarzschild geometry . When factor $`\frac{Q_{}q}{\sigma }`$ is positive potential varies smoothly. When $`\frac{Q_{}q}{\sigma }`$ becomes negative $`V_\pm `$ diverges at a certain location $`r=\alpha `$. For the second case factor $`\left(1+\frac{Q_{}q}{(rr_+)\sigma }\right)`$ vanishes at $`r=r_+\frac{Q_{}q}{\sigma }>r_+`$ and then becomes negative. At $`r=\alpha >r_+`$ denominator of $`V_\pm `$ vanishes. For all other cases $`\alpha <r_+`$ always, so there is no scope to diverge the potential. Thus for the positive energy solution when the electro-magnetic scalar potential in the field is of attractive nature corresponding potential diverges again for negative energy solution potential diverges for repulsive electro-magnetic scalar potential. For the integral spin particle, it is found that when potential diverges energy extraction is possible i.e., super-radiation is occurred in the space-time . On the other hand for the case of spin-half particle in Kerr geometry although at a certain parameter region potential diverges but super-radiation does not exist . In the case of spherically symmetric Schwarzschild geometry potential does not diverge at all and no scope of super-radiation . Here it is interesting to note that although our space-time is spherically symmetric but due to presence of electromagnetic interaction term, the region exist which is expected to be super-radiant.
Figure 1 shows behaviour of potential $`V_+`$ for different values black hole charges, where $`\sigma =0.8`$, $`m_p=0.8`$, $`l=\frac{1}{2}`$, $`q=1`$ are chosen; $`\alpha <r_+`$. When $`Q_{}=0`$ (solid curve), potential reduces to same as Schwarzschild case shown in Fig. 2 by Mukhopadhyay & Chakrabarti . It is also seen that with the increment of charge of the black hole, barrier height decreases. Increment of black hole charge indicates the increment of electro-magnetic coupling and corresponding repulsive scalar potential opposes the attractive gravitational field. So net effect decreases. Figure 2 shows the change of potential barrier for different values of particle charge, where $`\sigma =0.8`$, $`m_p=0.8`$, $`l=\frac{1}{2}`$, $`Q_{}=0.6`$ are chosen; $`\alpha <r_+`$. Solid curve indicates the potential felt by the neutron like particle.
Now come to the cases when $`\frac{Q_{}q}{\sigma }`$ is negative. For these cases $`\widehat{r}_{}r`$ relation attends multivalues. For both $`r\mathrm{}`$ and $`rr_+`$, $`\widehat{r}_{}\mathrm{}`$. As explained above net potential barrier diverges at a certain location in this parameter region. From equation (20) it is very clear that near $`r=\alpha `$, potential varies as $`\frac{1}{(r\alpha )^3}`$. So it has two branches, one repulsive and another attractive on each side of the singular point. As a result super-radiation is absent for the case of Reissner-Nordstrรถm geometry as in other cases . We can choose any combination of $`Q_{}`$, $`q`$ and $`\sigma `$ in such a way that $`\frac{Q_{}q}{\sigma }`$ is negative.
In Fig. 3 we show how nature of the potential ($`V_+`$) changes with rest mass of the incoming particle where, $`\sigma =0.8`$, $`Q_{}=0.5`$, $`l=\frac{1}{2}`$, $`q=1`$ are chosen. Solid curve shows nature for neutrino wave. It is very clear from the figure that with the increase of rest mass of the incoming particle gravitational interaction increases and corresponding potential barrier attains high value.
## IV. The complete solution
Now we will find spatially complete solution. As we mentioned earlier that solution of the angular part is known which is same as Schwarzschild case . For radial solution we need to solve decoupled radial equation. The solution of equation (16) for potential $`V_+`$ and $`V_{}`$, using Instantaneous WKB Approximation (in short IWKB) method can be written as
$$Z_+=\sqrt{T_+[k_+(\widehat{r}_{})]}exp(iu_+)+\sqrt{R_+[k_+(\widehat{r}_{})]}exp(iu_+),$$
$`(21a)`$
$$Z_{}=\sqrt{T_{}[k_{}(\widehat{r}_{})]}exp(iu_{})+\sqrt{R_{}[k_{}(\widehat{r}_{})]}exp(iu_{}),$$
$`(21b)`$
where,
$$k_\pm (\widehat{r}_{})=\sqrt{\left(\sigma ^2V_\pm \right)},$$
$`(22)`$
$$u_\pm (\widehat{r}_{})=k_\pm (\widehat{r}_{})๐\widehat{r}_{}+\mathrm{constant},$$
$`(23)`$
with
$$T_+(r)+R_+(r)=1,T_{}(r)+R_{}(r)=1\mathrm{i}\mathrm{n}\mathrm{s}\mathrm{t}\mathrm{a}\mathrm{n}\mathrm{t}\mathrm{a}\mathrm{n}\mathrm{e}\mathrm{o}\mathrm{u}\mathrm{s}\mathrm{l}\mathrm{y}.$$
$`(24)`$
Here, $`k`$ is the wavenumber of the incoming wave and $`u`$ is the Eiconal, $`T_\pm `$ and $`R_\pm `$ are instantaneous transmission and reflection coefficients respectively. Using this method at each location, instantaneously, WKB method is applied. This solution is valid when $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$, otherwise different method should be used.
In Fig. 4, the comparison of instantaneous reflection and transmission coefficients in between Schwarzschild and Reissner-Nordstrรถm geometry are shown. The parameters chosen are given in Figure Caption. With the decrease of barrier height the transmission coefficient increases as well as reflection coefficient decreases. As it is seen that by introduction of the electromagnetic coupling, potential barrier heights reduce so corresponding transmission probability increase with respect to that of Schwarzschild case (the behaviour for Schwarzschild case is graphically shown in ) for a particular set of parameter. So the presence of the charge of black hole decreases the curved nature of space-time.
Now recombining $`Z_+`$ and $`Z_{}`$ one easily can find out original radial Dirac wave functions $`R_{\frac{1}{2}}`$ and $`R_{\frac{1}{2}}`$ . Finally we will have complete solution as $`J(r,\theta )=R_{\pm \frac{1}{2}}(r)S_{\pm \frac{1}{2}}(\theta )`$.
## V. conclusions
In this paper, we have studied analytically the scattering of spin-half particles off Reissner-Nordstrรถm black hole. Our main motivation is to show analytically how the spin-half particles behave in the presence of electromagnetic interaction in curved space-time. We introduced the gravitational and electromagnetic gauge fields. Since no such kind of study had been carried out previously we started from scratch. Firstly, we wrote corresponding dynamical equation of spin-half particle namely Dirac equation in combined gravitational and electromagnetic background. Due to curvature of the space-time gravitational gauge field (here, spin coefficients for Reissner-Nordstrรถm geometry) was introduced. The electromagnetic interaction comes into the game because of charge of the black hole. Here, we have considered steady-state problem and corresponding components of electromagnetic vector potential to zero. We then separated the equation into radial and angular parts. It is seen that in case of spherically symmetric space-time, presence of charge of the gravitating object does not affect the behaviour of the incoming particles in polar direction. Only the radial part of the equation is influenced. We then decoupled the radial Dirac equation. Now the potential is dependent on charge-charge coupling in the space-time. If the charge of the black hole reduces to zero, the potential reduces to that of Schwarzschild case. With the presence of repulsive (or attractive) charge-charge interaction for positive (or negative) energy solution magnitude of curvature effect reduces. This is because of opposing nature of two simultaneous interactions.
There is one interesting sector of the solution (which sector was absent in uncharged spherically symmetric space-time). If the charge-charge interaction is of attractive nature for positive energy solution (or repulsive for negative energy solution) then potential at a certain location ($`r=\alpha `$) diverges. But because of $`\frac{1}{(r\alpha )^3}`$ variance of potential super-radiation is absent.
Here we study the behaviour of potential by varying charge of the black hole, charge of the incoming particle, rest mass of the incoming particle. We also study the space-dependent reflection and transmission coefficients and show graphically for one set of physical parameter. It is seen that as potential barrier height decreases corresponding transmission probability increases. We solve the radial Dirac equation by IWKB method.
## Acknowledgment
I am thankful to Prof. Sandip K. Chakrabarti regarding selection of this problem. I also like to thank to Kaushik Ghosh for helpful comments during my work.
## figure captions
Fig. 1: Behaviour of potential for different values of the black hole charge. Fixed parameters are, $`\sigma =0.8`$, $`m_p=0.8`$, $`l=\frac{1}{2}`$ and $`q=1`$. From upper to lower curves the charge $`Q_{}`$ of the black holes are chosen as $`0,0.2,0.4,0.6,0.8,0.998`$.
Fig. 2: Behaviour of potential for different values of the incoming particle charge. Fixed parameters are, $`\sigma =0.8`$, $`m_p=0.8`$, $`l=\frac{1}{2}`$ and $`Q_{}=0.6`$. From upper to lower curves the charge $`q`$ of the particles are chosen as $`0,0.2,0.4,0.6,0.8,1`$.
Fig. 3: Behaviour of potential for different values of the rest mass of incoming particle. Fixed parameters are, $`\sigma =0.8`$, $`Q_{}=0.5`$, $`l=\frac{1}{2}`$ and $`q=1`$. From upper to lower curves the mass $`m_p`$ of the particles are chosen as $`0.4,0.3,0.2,0.1,0`$.
Fig. 4: Instantaneous reflection (R) and transmission (T) coefficients for Reissner-Nordstrรถm (solid curves) and Schwarzschild (dotted curves) black holes. Physical parameters are chosen as $`\sigma =0.8,m_p=0.8,l=\frac{1}{2},q=1`$. For Reissner-Nordstrรถm case $`Q_{}=0.5`$. |
warning/0003/cond-mat0003211.html | ar5iv | text | # Unconventional odd-parity superconductivity in the ladder compounds
\[
## Abstract
Superconductive properties of the two-leg ladder compounds are studied theoretically. The antiferromagnetic fluctuations are considered because of the good nesting of the Fermi surfaces. The attractive interaction which is most likely due to the electron-phonon coupling is also taken into account. Under this circumstance, it is shown that the superconductivity has two sets of spin-triplet pairings. The gaps for both pairings have no node.
\]
A number of unconventional superconductivities are under intensive study in recent years. Examples include the high $`T_\mathrm{c}`$ cuprates, heavy fermions, Sr<sub>2</sub>RuO<sub>4</sub>, organic conductors etc. The common property of these compounds is the existence of antiferromagnetic(AF) fluctuations. Because of this a large number of works are devoted to understand the unconventional nature from AF fluctuations. As a result most of them neglect the electron-phonon interactions which explain most of the superconductors before the discovery of the high $`T_\mathrm{c}`$ cuprates. Although, there exist a few attempts to explain the unconventional nature of superconductivity from attractive interactions (which most likely come from electron-phonon coupling) under the influence of AF fluctuations.
Recently the ladder compounds which may have some connections with the high $`T_\mathrm{c}`$ cuprates were found. The crystal structures are different but the basic role of Cu and O looks similar. Hiroi et al. were the first to synthesize the family of layer compounds Sr<sub>n-1</sub>Cu<sub>n</sub>O<sub>2n-1</sub>, which have arrays of parallel line defects. Nearly ideal ladder compounds should result. The first member ($`n=2`$ or SrCu<sub>2</sub>O<sub>3</sub>) has two-leg ladders, the second ($`n=3`$ or Sr<sub>2</sub>Cu<sub>3</sub>O<sub>5</sub>) has three-leg ladders and so on. It was also clearly shown that the material LaCuO<sub>2.5</sub> has an insulator-metal transition upon hole doping by substitution of Sr<sup>2+</sup> for La<sup>3+</sup> but no sign of superconductivity was observed. There have been considerable interests in magnetism of the ladder compounds .
The ladder material (Sr, Ca)<sub>14</sub>Cu<sub>24</sub>O<sub>41-ฮด</sub> was synthesized by MaCarron et al. and by Siegrist et al.. In this compound superconductivity was found by Uehara et al.. It appears at around $`T_\mathrm{c}10`$K but only under high pressure more than 3GPa. Due to this limitation essential properties of these compounds have not yet clarified by experiments. In particular the nature of superconductivity is not well understood.
In this paper we study the superconductive properties of two-leg(two-chain) ladders theoretically. In order to obtain universal features we only consider the minimum model which is supposed to give essential features of superconductivity. Details which are pertinent to specific compounds are not considered. Neither did we try to estimate the values of physical quantities like $`T_\mathrm{c}`$. This is because, for quantitative predictions, one needs to fix various physical quantities which are not known theoretically or experimentally. Although, as shown below, $`T_\mathrm{c}`$ is similar to those of the classical BCS superconductors.
The existence of unconventional spin-triplet superconductivity in the ladder systems is shown. This result seems to be robust and there is a possibility to find unconventional superconductivity in other ladder materials which include more than two legs ladders.
Band structure
The energy dispersion of the noninteracting two-chain ladder is given
$`\epsilon (k)=2t\mathrm{cos}k_x\pm t^{},`$ (1)
where $`t`$ is the transfer integral in the $`x`$-direction which is along the chains (usually denoted as $`c`$-axis direction), and $`t^{}`$ is the transfer integral between the two chains. The two transfer integrals, $`t`$ and $`t^{}`$ have the same order of magnitude. The sign in front of $`t^{}`$ represents parity with respect to exchange of the two chains.
The inter-ladder hoppings are small and estimated to be $`t^{\prime \prime }1/20\mathrm{to}1/30t^{}`$, and we shall neglect them hereafter.
There are four Fermi surfaces as shown in Fig. 1. The superconductive pairs with finite momenta are in general have smaller binding energies compared with those with momentum zero, i.e. those consist of electrons in the same band (either $`+`$ or $``$ in Eq.(1)). Therefore we shall consider only one band and the two Fermi surfaces for the moment.
Attractive interaction
First let us study interactions which are attractive and do not change spins. See Fig. 2. The general form is
$`H_{\mathrm{int}}={\displaystyle \underset{k,k^{}}{}}{\displaystyle \underset{\alpha ,\beta }{}}:f(q)a_{k+q,\alpha }^{}a_{k,\alpha }a_{k^{}q,\beta }^{}a_{k^{},\beta }:,`$ (2)
where $`f(q)>0`$ represents an attractive interaction and mainly depends on $`q_x`$. It has a peak at $`q_x=0`$. Here $`::`$ denotes creation-annihilation normal ordering. This interaction most likely comes from electron-phonon coupling. Let us concentrate on the interactions which are relevant to $`(k,k)`$ pairing. Thus choose $`k^{}=k`$ and put $`q=k^{\prime \prime }k`$
$`H_{\mathrm{int}}={\displaystyle \underset{k,k^{\prime \prime }}{}}{\displaystyle \underset{\alpha ,\beta }{}}f(k^{\prime \prime }k)a_{k^{\prime \prime },\alpha }^{}a_{k^{\prime \prime },\beta }^{}a_{k,\alpha }a_{k,\beta }.`$ (3)
(4)
Let us introduce $`V_{s_1s_2s_3s_4}(k,k^{})`$ by
$`H_{\mathrm{int}}`$ (5)
$`={\displaystyle \frac{1}{2}}{\displaystyle \underset{k,k^{}}{}}{\displaystyle \underset{s_1,s_2,s_3,s_4}{}}V_{s_1s_2s_3s_4}(k,k^{})a_{k,s_1}^{}a_{k,s_2}^{}a_{k^{},s_3}a_{k^{},s_4},`$ (6)
where $`s_i^{}s`$ are spin indices. By the symmetry
$`V_{s_1s_2s_3s_4}(k,k^{})`$ $`=`$ $`V_{s_2s_1s_3s_4}(k,k^{})`$ (7)
$`=`$ $`V_{s_1s_2s_4s_3}(k,k^{}),`$ (8)
we obtain
$`V_{s_1s_2s_3s_4}(k,k^{})`$ (9)
$`={\displaystyle \frac{1}{2}}\{f(k+k^{})+f(kk^{}))\}\delta _{s_1s_3}\delta _{s_2s_4}`$ (10)
$`{\displaystyle \frac{1}{2}}\{f(kk^{})+f(k+k^{})\}\delta _{s_1s_4}\delta _{s_2s_3}.`$ (11)
Magnetic interaction
The general form of magnetic interaction is written (see Fig. 3)
$`H_{\mathrm{int}}^\mathrm{m}`$ (12)
$`={\displaystyle \underset{k,k^{},q}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}:J_x(q)(\sigma _x)_{\alpha \beta }(\sigma _x)_{\gamma \delta }a_{k+q,\alpha }^{}a_{k,\beta }a_{k^{}q,\gamma }^{}a_{k^{},\delta }:`$ (13)
$`{\displaystyle \underset{k,k^{},q}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}:J_y(q)(\sigma _y)_{\alpha \beta }(\sigma _y)_{\gamma \delta }a_{k+q,\alpha }^{}a_{k,\beta }a_{k^{}q,\gamma }^{}a_{k^{},\delta }:`$ (14)
$`{\displaystyle \underset{k,k^{},q}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}:J_z(q)(\sigma _z)_{\alpha \beta }(\sigma _z)_{\gamma \delta }a_{k+q,\alpha }^{}a_{k,\beta }a_{k^{}q,\gamma }^{}a_{k^{},\delta }:.`$ (15)
(16)
In general $`J_i(q)(>0)`$โs ($`i=x,y,z`$) are different each other, but for simplicity we treat explicitly only the case $`J_x(q)=J_y(q)=J_z(q)J(q)`$ in the following. The anisotropy of the magnetic interaction amplifies the tendency toward the odd-parity superconductivity, and it determines the direction of the $`\stackrel{}{d}`$ vector ( which is defined by Eq.(LABEL:dvec)) . If the anisotropy is strong enough, the odd-parity superconductivity is realized even if $`f(q)=0`$.
By considering only the terms which contribute ($`k,k`$) pairings, we have (by putting $`k=k^{}`$)
$`H_{\mathrm{int}}^\mathrm{m}`$ (17)
$`={\displaystyle \underset{k,q}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}:J(q)(\stackrel{}{\sigma })_{\alpha \beta }(\stackrel{}{\sigma })_{\gamma \delta }a_{k+q,\alpha }^{}a_{k,\beta }a_{kq,\gamma }^{}a_{k,\delta }:.`$ (18)
Put $`q=k^{\prime \prime }k`$ then
$`H_{\mathrm{int}}^\mathrm{m}`$ (20)
$`={\displaystyle \underset{k,k^{\prime \prime }}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}:J(k^{\prime \prime }k)(\stackrel{}{\sigma })_{\alpha \beta }(\stackrel{}{\sigma })_{\gamma \delta }a_{k^{\prime \prime },\alpha }^{}a_{k,\beta }a_{k^{\prime \prime },\gamma }^{}a_{k,\delta }:`$ (21)
$`={\displaystyle \underset{k,k^{\prime \prime }}{}}{\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta }{}}J(k^{\prime \prime }k)(\stackrel{}{\sigma })_{\alpha \beta }(\stackrel{}{\sigma })_{\gamma \delta }a_{k^{\prime \prime },\alpha }^{}a_{k^{\prime \prime },\gamma }^{}a_{k,\beta }a_{k,\delta }.`$ (22)
(23)
By comparing this equation (23) and Eq.(6), and using the symmetry (8), the magnetic interaction is written
$`V_{s_1s_2s_3s_4}^\mathrm{m}(k,k^{})`$ (24)
$`={\displaystyle \frac{1}{2}}\{J(k+k^{})+J(kk^{})\}(\stackrel{}{\sigma })_{s_1s_3}(\stackrel{}{\sigma })_{s_2s_4}`$ (25)
$`{\displaystyle \frac{1}{2}}\{J(kk^{})+J(k+k^{})\}(\stackrel{}{\sigma })_{s_1s_4}(\stackrel{}{\sigma })_{s_2s_3}.`$ (26)
Gap equation
We use the weak coupling BCS theory. It is possible that the strong-coupling corrections modify the results quantitatively but not qualitatively because $`T_\mathrm{c}`$ is not so high. The gap equation is
$`\mathrm{\Delta }_{ss^{}}(k)=`$ $``$ $`{\displaystyle \underset{k^{},s_3,s_4}{}}V_{s^{}ss_3s_4}(k,k^{})`$ (27)
$`\times `$ $`{\displaystyle \frac{\mathrm{\Delta }_{s_3s_4}(k^{})}{2E_k^{}}}\mathrm{tanh}\left({\displaystyle \frac{\beta E_k^{}}{2}}\right),`$ (28)
where the (scalar) excitation energy $`E_k^{}`$ is given by
$`E_k=\sqrt{\epsilon (k)^2+{\displaystyle \frac{1}{2}}\mathrm{tr}(\mathrm{\Delta }(k)\mathrm{\Delta }(k)^{})},`$ (29)
For even-parity states (spin-singlet pairing) one can write
$`\mathrm{\Delta }(k)`$ $`=`$ $`i\sigma _y\psi (k)`$ (30)
with $`\psi (k)=\psi (k)`$. Thus the gap equation for even-parity states is
$`\psi (k)={\displaystyle \underset{k^{}}{}}V_{kk^{}}^{\mathrm{even}}{\displaystyle \frac{\psi (k^{})}{2E_k^{}}}\mathrm{tanh}\left({\displaystyle \frac{\beta E_k^{}}{2}}\right).`$ (31)
where $`V_{kk^{}}^{\mathrm{even}}`$ is obtained from Eq.(11) and Eq.(26) as
$`V_{kk^{}}^{\mathrm{even}}`$ (34)
$`={\displaystyle \frac{1}{2}}\{f(k+k^{})+f(kk^{})+f(kk^{})+f(k+k^{})\}`$
$`+{\displaystyle \frac{3}{2}}\{J(k+k^{})+J(kk^{})+J(kk^{})+J(k+k^{})\}.`$
(35)
For odd-party states (spin-triplet pairing) one can write
$`\mathrm{\Delta }(`$ $`k`$ $`)=i(\stackrel{}{d}(k)\stackrel{}{\sigma })\sigma _y`$ (36)
with $`\stackrel{}{d}(k)=\stackrel{}{d}(k)`$. The gap equation for odd-parity states is
$`\stackrel{}{d}(k)`$ $`={\displaystyle \underset{k^{}}{}}V_{kk^{}}^{\mathrm{odd}}{\displaystyle \frac{\stackrel{}{d}(k^{})}{2E_k^{}}}\mathrm{tanh}\left({\displaystyle \frac{\beta E_k^{}}{2}}\right).`$ (38)
where $`V_{kk^{}}^{\mathrm{odd}}`$ is obtained from Eq.(11) and Eq.(26) as
$`V_{kk^{}}^{\mathrm{odd}}`$ (39)
$`={\displaystyle \frac{1}{2}}\{f(k+k^{})+f(kk^{})f(kk^{})f(k+k^{})\}`$ (40)
$`+{\displaystyle \frac{1}{2}}\{J(k+k^{})+J(kk^{})J(kk^{})J(k+k^{})\}.`$ (41)
(42)
BCS approximation
Following the BCS treatment of superconductivity we make further approximations. Namely the pairing interactions are constant within the cutoff and zero outside the cutoff. The function $`f(q)`$ has a peak at $`q_x=0`$ but it is approximated by a constant $`V>0`$. Therefore this interaction corresponds to the phonon-mediated interaction in the classical BCS theory. In addition to this interaction the ladder compounds have interactions due to AF fluctuations which are induced by nesting of the Fermi surfaces in a similar fashion to some of the organic conductors or the high $`T_\mathrm{c}`$ cuprates in the low doping region. Since the Fermi surfaces are straight and parallel, the nesting is perfect. Therefore the nesting vector is $`Q=(c,k_y,k_z)`$ where c is the distance between the two Fermi surfaces of the same band. See Fig.1. (Nesting is also perfect between the Fermi surfaces of different bands. But these are not relevant to the superconducting pairing of $`(k,k)`$.) Therefore there exist AF fluctuations with nesting vector $`Q`$. Now we take the magnetic interaction $`J(q)`$ to be constant $`J(Q)>0`$ when $`q`$ is is close to $`Q`$ and connects two points in different Fermi surfaces (of the same band) within the cutoff. Then Eq.(35) gives
$`V^{\mathrm{even}}=V+3J(Q)`$ (43)
for even-parity states, and Eq.(42) gives
$`V^{\mathrm{odd}}=V+J(Q)`$ (44)
for odd-parity states. Therefore we always have $`V^{\mathrm{odd}}>V^{\mathrm{even}}`$ as long as magnetic interactions are present. It leads to odd-parity (spin-triplet) pair condensation with
$`T_\mathrm{c}=1.13\omega _\mathrm{D}e^{1/\{N(k_\mathrm{F})V^{\mathrm{odd}}\}}.`$ (45)
where $`\omega _\mathrm{D}`$ is a cutoff.
When the anisotropy of AF fluctuations is taken into account, $`V^{\mathrm{odd}}`$ is replaced by
$`V^{\mathrm{odd}}=\{\begin{array}{cc}VJ_x(Q)+J_y(Q)+J_z(Q)\hfill & \text{for }d_x0\hfill \\ VJ_y(Q)+J_z(Q)+J_x(Q)\hfill & \text{for }d_y0\hfill \\ VJ_z(Q)+J_x(Q)+J_y(Q)\hfill & \text{for }d_z0\hfill \end{array}.`$ (49)
It can be seen from Eq.(49) that if AF fluctuations perpendicular to the $`\stackrel{}{d}`$ vector are strong enough, there is no superconductive condensation. It also can be seen that AF fluctuations parallel to the $`\stackrel{}{d}`$ vector promote the odd-parity superconductivity. It is rather straightforward to show that there is no node in the gaps from the gap equation (28).
Discussions
To conclude, we study the unconventional superconductive nature of the two-leg ladder compounds. The attractive interactions which are most likely due to the electron-phonon coupling are considered. AF fluctuations from the good nestings of the Fermi surfaces are also included.
Our main conclusion is that the two-leg ladder compounds have two sets of spin-triplet superconductive states without nodes in the gap corresponding the two sets of the Fermi surfaces. These superconductivities are suppressed by AF fluctuations perpendicular to the $`\stackrel{}{d}`$ vector and vanish if the AF fluctuations are strong enough. Our results also apply to the ladder compound with more than two legs.
It is possible to have two $`T_\mathrm{c}`$โs corresponding to the two bands, respectively. However it is likely to have a single $`T_\mathrm{c}`$ due to some proximity effects induced by the interactions which are not included here.
Acknowledgments
It is a pleasure to thank J. Akimitsu and N. Mori for useful discussions. |
warning/0003/quant-ph0003074.html | ar5iv | text | # On the nature of continuous physical quantities in classical and quantum mechanics
## 1 Introduction
Consider the following experimental setup: We have a source of quantum-mechanical particles (e.g. electrons), a fluorescent screen, and a barrier between the source and the screen which has one or more openings. The barrier, with its openings, prepares an ensemble of particles whose state, upon arrival at the screen, is given by some wavefunction $`\psi `$. Letโs assume for simplicity that the screen is one-dimensional and infinite in both directions (i.e., represented by $``$). The screen should be thought of as a measuring device for the position of particles in the ensemble, and we know that the pattern of shading on the screen will be distributed in accordance with the squared modulus $`|\psi |^2`$ of the wavefunction. For any individual particle in the ensemble, and any region $`S`$ on the screen, $`_S|\psi |^2`$ gives the probability that the the screen will flash inside $`S`$ when the particle hits the screen. It would be very natural then to attempt to interpret the probability distribution $`|\psi |^2`$ as a measure of our ignorance about the precise state of individual particles in the ensemble. That is, it is natural to think that each particle in the ensemble does in fact have a โsure-fireโ (probability $`1`$) disposition to be found at a certain point $`\lambda `$, and probabilities arise out of our ignorance about precisely which point $`\lambda `$ that is.
It is clear that the traditional interpretive obstacles in quantum theory do not prevent us from thinking this way of this measurement. For one, since we are not asking how a particle gets from the source to the screen, interference effects play no role. Second, since we are in a position measurement context (i.e., we are ignoring other observables such as momentum), there can be no question of the no-hidden-variables theorems presenting any obstacles to our thinking of $`|\psi |^2`$ as a mixture of pure position states.
However, there remains a serious difficulty for supplying an ignorance interpretation of $`|\psi |^2`$: There does not appear to be any object in the standard Hilbert space formalism of QM which could represent the state of an individual particle in the ensemble. According to Teller , then, attempting to give an ignorance intepretation to $`|\psi |^2`$ amounts to questioning the โdescriptive completenessโ of quantum theory:
> โโฆif imprecise or imperfectly defined values of a continuous quantity are to be explained in terms of assumed underlying states with exact point values, these assumed underlying states cannot be described within the theory, and the theory is descriptively incomplete.โ (p. 347)
But since QM is descriptively complete, $`|\psi |^2`$ should not be interpreted as a measure of our ignorance of the precise state of the individual particles.
Now it is certainly true that no *vector* in Hilbert space can be taken to describe the state of an individual particle with a sharp location. This is just the familiar point that operators with continuous spectra, like position, have no eigenvectors. Furthermore, Gleasonโs Theorem entails that any (pure) $`\sigma `$-additive probability measure on the logic of quantum propositions (i.e., lattice of subspaces of a Hilbert space) is given by a vector $`\psi `$, via the standard formula $`E\psi |E\psi `$. I believe, however, that the $`\sigma `$-additivity assumption in Gleasonโs theorem is too restrictive. Indeed, I will argue that if we required $`\sigma `$-additivity of the states of a *classical* mechanical system, then we would similarly have to conclude โ in direct contradiction to what we know to be true โ that there can be no states assigning sharp values to continuous quantities in that case as well. In fact, I will show that in both cases it is possible to assign sharp values to continuous quantities using finitely-additive (โsingularโ) states. I will also demonstrate that difficulties which arise when using singular states in conjunction with the traditional von-Neumann account of ideal measurement are resolved when we adopt a more realistic account of measurement.
## 2 Observables, properties, and states
It is well-known that there are methods of extending the Hilbert space formalism of QM so that it includes โgeneralized eigenstatesโ for continuous spectrum observables (e.g., the rigged Hilbert space formalism). However, these constructions are not needed for my purposes: Any continuous spectrum observable is associated with a Boolean algebra of spectral projections, and a 2-valued homomorphism on this Boolean algebra gives a generalized eigenstate. Furthermore, such homomorphisms should not be considered any less a part of the Hilbert space formalism in the continuous case than they are in the discrete case, where they just happen to be represented by proper eigenvectors.
### 2.1 Preliminaries
I begin by establishing the terminology I will be using in this paper. Let $``$ be an arbitrary $`\sigma `$-complete Boolean algebra, with additive identity $`\mathrm{๐}`$ and multiplicative identity $`๐`$. We say that $`\omega `$ is a *state* of $``$ just in case $`\omega `$ is a mapping of $``$ into $`[0,1]`$ such that $`\omega (๐)=1`$ and
$$\omega (AB)=\omega (A)+\omega (B),$$
(1)
for each disjoint $`A,B`$, i.e., $`AB=\mathrm{๐}`$. Clearly, the set of states on $``$ is a convex set. If $`\omega `$ is an extreme point among the states on $``$, we say that $`\omega `$ is a *pure* state. That is, $`\omega `$ is pure just in case: If $`\omega =a\rho +(1a)\tau `$, for some $`a(0,1)`$ and states $`\rho ,\tau `$ on $``$, then $`\omega =\rho =\tau `$. The state $`\omega `$ is pure if and only if it is a Boolean homomorphism of $``$ onto $`\{0,1\}`$. If $`\omega `$ is not pure, we say that $`\omega `$ is *mixed*. We say that a state $`\omega `$ is *normal* just in case
$$\omega \left(_{i=1}^{\mathrm{}}A_i\right)=\underset{i=1}{\overset{\mathrm{}}{}}\omega (A_i),$$
(2)
for each countable pairwise disjoint sequence $`\{A_i\}`$ in $``$. If $`\omega `$ is not normal, we say that $`\omega `$ is *singular*. Recall that a family $``$ of elements in $``$ is called a *filter* just in case: (i) $`\mathrm{}`$, (ii) $`AB`$, when $`A,B`$, and (iii) if $`A`$ and $`AB`$, then $`B`$. If, furthermore, either $`A`$ or $`\neg A`$ for all $`A`$, then $``$ is said to be an *ultrafilter*. Each ultrafilter $`๐ฐ`$ on a Boolean algebra $``$ is the preimage of $`1`$ under some pure state $`\omega `$ of $``$, and $`\omega `$ is normal if and only if $`๐ฐ`$ is closed under countable meets. For any non-zero element $`A`$, the set $`๐ฐ_A=\{B:AB\}`$ is a filter, and is an ultrafilter if $`A`$ is an atom in $``$. In this case, $`๐ฐ_A`$ is called the *principal ultrafilter* generated by $`A`$. It is easy to see then that principal ultrafilters give rise to normal pure states.
The discussion of this paper focuses on two Boolean algebras: The $`\sigma `$-complete Boolean algebra $`()`$ of Borel subsets of $``$, and its quotient $`()/๐ฉ`$ by the ideal of Lebesgue measure zero sets. Recall that $`()`$ is the (Boolean) $`\sigma `$-algebra generated by the open subsets of $``$. The logical operations on $`()`$ are just the set-theoretic operations $``$ (intersection), $``$ (union), and $``$ (complement in $``$). Clearly, $`()`$ is purely atomic; i.e., for each $`S()`$ there is some atomic element $`A()`$ (i.e., a singleton set) such that $`AS`$. Thus, by the considerations adduced above, $`()`$ has an abundance of pure normal states.
Let $`๐ฉ`$ denote the family of Lebesgue measure zero sets in $`()`$. We define an equivalence relation $``$ on $`()`$ by setting $`S_1S_2`$ just in case $`S_1\mathrm{}S_2๐ฉ`$, where $`S_1\mathrm{}S_2=(S_1\backslash S_2)(S_2\backslash S_1)`$. For each $`S()`$, let $`\pi (S)=\{S^{}():S^{}S\}`$, and let
$$()/๐ฉ=\{\pi (S):S()\}.$$
(3)
For any countable family $`\{\pi (S_i)\}`$ of elements in $`()/๐ฉ`$, we define,
$$_{i=1}^{\mathrm{}}\pi (S_i)=\pi \left(_{i=1}^{\mathrm{}}S_i\right).$$
(4)
(This is well-defined since $`๐ฉ`$ is a $`\sigma `$-ideal.) For $`\pi (S)()/๐ฉ`$, we set $`\neg \pi (S)=\pi (S)`$. It then follows that $`()/๐ฉ,,\neg `$ is a $`\sigma `$-complete Boolean algebra. It is also possible to show โ although I do not prove it here โ that $`()/๐ฉ`$ has *no* atoms; i.e., for each nonzero $`A()/๐ฉ`$, there is some nonzero $`B()/๐ฉ`$ such that $`B<A`$. This, in fact, is equivalent to the nonexistence of pure normal states for $`()/๐ฉ`$, which I do prove below in Proposition 1.
### 2.2 The position observable in QM
It is well-known that no wavefunctions $`\psi `$ in a Hilbert space give a probability density $`|\psi |^2`$ that is focused at one point. In this section, I will show that this well-known fact may be described in purely logical terms as the nonexistence of pure normal states on the Boolean algebra $`()/๐ฉ`$.
The state space of quantum mechanics is obtained by requiring an irreducible representation of the canonical commutation relation (CCR):
$$[\widehat{Q},\widehat{P}]=i\mathrm{}I,$$
(5)
by operators $`\widehat{Q},\widehat{P}`$ on some Hilbert space $``$. In its integrated (Weyl) form, there is a unique (up to unitary isomorphism) representation of the CCR on the Hilbert space $`L_2()`$ of equivalence classes of square integrable functions from $``$ into $``$ . (Two functions are equivalent just in case they agree except possibly on a set of measure zero.) In this representation, the position operator $`\widehat{Q}`$ is defined on a dense subset of $`L_2()`$ by the equation
$$(\widehat{Q}\psi )(q)=q\psi (q).$$
(6)
A โpropositionโ about the position of the particle is then given by a projection operator which is a function of $`\widehat{Q}`$. In particular, for each $`S()`$, we define an operator $`E(S)`$ on $`L_2()`$ by the equation
$$(E(S)\psi )(q)=\chi _S(q)\psi (q).$$
(7)
The set $`๐ช_Q=\{E(S):S()\}`$ is a $`\sigma `$-complete Boolean algebra of projection operators on the Hilbert space $``$, and $`E`$ is a Boolean $`\sigma `$-homomorphism from $`()`$ onto $`๐ช_Q`$. Moreover, $`E(S)=\mathrm{๐}`$ if and only if $`S`$ has Lebesgue measure zero. Thus, the kernel of the mapping $`E`$ is $`๐ฉ`$, and (by the first isomorphism theorem for rings) $`๐ช_Q`$ is isomorphic to the quotient Boolean algebra $`()/๐ฉ`$. ($`E`$ is usually called the spectral measure for $`\widehat{Q}`$.)
Now, every wavefunction $`\psi `$ gives rise to a normal state $`\rho _\psi `$ on $`๐ช_Q`$ in the standard way:
$$\rho _\psi (E(S)):=\psi |E(S)\psi =_{}\chi _S(q)|\psi (q)|^2๐q=_S|\psi (q)|^2๐q.$$
(8)
However, it follows from purely logical considerations that there cannot be normal โeigenstatesโ for $`\widehat{Q}`$.
###### Proposition 1.
There are no pure normal states of $`()/๐ฉ`$.
###### Proof.
Let $`\omega `$ be a pure state of $`()/๐ฉ`$, and let $`๐ฐ=\omega ^1(1)`$ be the corresponding ultrafilter. (Case 1) Suppose that $`\pi (S)๐ฐ`$ for all compact subsets $`S`$ of $``$. For each $`n`$, let $`A_n=[n,n]`$. Then, $`\pi (A_n)๐ฐ`$ and since $`๐ฐ`$ is an ultrafilter, $`\pi (A_n)=\neg \pi (A_n)๐ฐ`$. However,
$$_{n=1}^{\mathrm{}}\pi (A_n)=\pi \left(_{n=1}^{\mathrm{}}A_n\right)=\mathrm{๐},$$
(9)
and therefore $`๐ฐ`$ is not closed under countable meets. (Case 2) If $`\pi (S)๐ฐ`$ for some compact set $`S`$ in $``$, then $`๐ฐ`$ contains $`\pi (_\lambda )`$ where $`_\lambda `$ is the family of open neighborhoods of some point $`\lambda `$. For each $`n`$, let
$$B_n=(\lambda n^1,\lambda +n^1).$$
(10)
Then, $`\pi (B_n)๐ฐ`$ for each $`n`$. However,
$$_{n=1}^{\mathrm{}}\pi (B_n)=\pi \left(_{n=1}^{\mathrm{}}B_n\right)=\pi \left(\{\lambda \}\right)=\mathrm{๐},$$
(11)
since $`\{\lambda \}`$ has Lebesgue measure zero. Thus, $`_{n=1}^{\mathrm{}}\pi (B_n)๐ฐ`$ and $`๐ฐ`$ is again not closed under countable meets. โ
Thus, assuming that quantum theory is descriptively complete, and that quantum states must be $`\sigma `$-additive, $`|\psi |^2`$ can not be interpreted as a measure of our ignorance of the true (pure) state of individual particles in the ensemble. Some philosophers draw the moral from this that quantum theory is teaching us something new about the nature of continuous quantities, namely that we should not think or talk about continuous quantities possessing point values, nor should we think or talk about statistical states as distributions over point-valued states. For example, Fine argues that we should think of statistical state $`\rho _\psi `$ as a map that assigns a *set*, rather than a numerical value, to $`\widehat{Q}`$. In particular,
$$\rho _\psi (\widehat{Q}):=\{S:\rho _\psi (E(S))=1\}.$$
(12)
Of course, $`\rho _\psi (\widehat{Q})`$ will never be a singleton, and โ perhaps contrary to our intuitions โ the statement that the particle lies in $`S`$ is a maximally specific description of the location of the particle.
Teller agrees with Fine that a statistical state should be interpreted not as a mixture of value states, but as assigning a โpartlessโ set value. Nonetheless, Teller does recognize that this description of the location of the particle can be misleading: it suggests โ contrary to the predictions of quantum theory โ that no matter where we look in $`S`$, we will find a โpieceโ of the particle. Teller suggests, however, that we should think of the property โbeing located in $`S`$โ as including,
> โa collection of dispositions or potentialities to manifest a more refined position of the same nature whenever a more refined measurement interaction takes place.โ (p. 357)
Thus, Teller seems to be claiming that the probability distribution $`|\psi |^2`$ itself gives a maximally specific description of an individual element in the ensemble.
Teller claims, moreover, that this dual property/disposition nature of position in QM is โnothing exceptional;โ in fact, it is just like the color white which is โa manifest property and at the same time an array of dispositionsโ (ibid). This analogy to color, however, seems to show exactly what leaves us uneasy about Tellerโs proposal. In the case of the color white, we can explain an objectโs dispositions to appear certain ways in certain contexts in terms of categorical properties of the object (viz., the chemical composition of its surface). On the other hand, Teller is claiming that a quantum-mechanical particle can have irreducibly probabilistic dispositions with no categorical basis (apart from, perhaps, $`\psi `$ itself, which is difficult to see as anything other than a catalog of those dispositions).
### 2.3 The ignorance interpretation in classical mechanics
According to Teller, these irreducibily probabilistic dispositions are a peculiar feature of quantum mechanics. Unlike the quantum case, the probabilities in classical mechanics may be thought of as measures of our ignorance of the precise state:
> โIn the context of classical physics the outcome of an inexact measurement may be described as a probability distribution over exact values, and we know how to state what this means: There is a probability (interpreted as a relative frequency, subjective degree of belief, or propensity) for the value of the measured quantity to have one or another of the exact values in the support of the distribution.โ (p. 352)
Let us attempt, then, to give a mathematically rigorous description of this purported distinction between the classical and quantum cases.
A classical particle moving in one dimension has a *phase space* $`^2`$, with one coordinate for the position $`Q`$ of the particle, and the other coordinate for the momentum $`P=m(dQ/dt)`$ of the particle. We may think of Borel subsets of $`^2`$ as representing statements that ascribe properties to the particle. (For background material, see Chap. 1 of Ref. References.) A dynamical variable is represented by a measurable, real-valued function $`F`$ on phase space. For example, the position of the particle is represented by the function $`F((q,p))=q`$. For any Borel set $`S`$ in $``$, $`F^1(S)`$ consists precisely of those โstatesโ (i.e., points in phase space) for which the value of $`F`$ lies in $`S`$. $`F^1`$ (considered as a set mapping) is in fact a Boolean $`\sigma `$-homomorphism from $`()`$ onto some Boolean subalgebra $`_F`$ of $`(^2)`$. For example, the disjunctive proposition *โThe value of $`F`$ lies either in $`S_1`$ or in $`S_2`$* is represented formally by
$$F^1(S_1S_2)=F^1(S_1)F^1(S_2).$$
(13)
Consider now the completely classical example of a person (or machine) Bob throwing darts at a (one-dimensional) dartboard. Since we are ignorant of the factors that influence each individual throw, we assign a probability density $`\rho `$ to the points on the dartboard (represented here by $``$). Now, just as in the quantum-mechanical case, the distribution $`\rho `$ on $``$ defines a mixed normal state on $`_F`$ by means of the equation
$$\rho (F^1(S))=_{}\chi _S(q)\rho (q)๐q=_S\rho (q)๐q.$$
(14)
Any particular dart in the โensembleโ should, however, be characterized by a *pure* state on $`_F`$. (i.e., there is some point $`\lambda `$ such that the dart has probability $`1`$ of hitting that point). And, in fact, everything works out fine in this case since there *are* pure normal states of $`_F`$. Indeed, the atoms in $`_F`$ are just lines in $`^2`$ on which $`F`$ is constant: $`A_\lambda =\{(q,p):F((q,p))=\lambda \}`$. Since each atom in a Boolean algebra gives rise to a pure normal state, there is a normal pure state $`\omega _\lambda `$ of $`_F`$ defined explicitly by the condition that $`\omega _\lambda (S)=1`$ if and only if $`A_\lambda S`$. Conversely, if $`\omega `$ is a pure normal state of $`_F`$, then $`\omega =\omega _\lambda `$ for some $`\lambda `$. \[First show that $`\omega (F^1(S))=1`$ for some compact set $`S`$. Standard topological arguments then show that $`\omega (F^1(V))=1`$ for every open neighborhood $`V`$ or some point $`\lambda `$. Now use the fact that $`\{\lambda \}`$ is the intersection of a countable family of open sets.\] Thus, the pure normal states on $`_F`$ are in one-to-one correspondence with the sets on which $`F`$ takes a constant value. Finally, it is easy to see that the distribution $`\rho `$ does admit interpretation as a measure of ignorance over the pure normal states, namely $`\rho =\omega _\lambda ๐\rho `$.
However, before we grant that there is a fundamental difference here between the quantum and classical cases, letโs think a bit more carefully about the interpretation of the mathematical objects weโve been dealing with. In particular, an element $`F^1(S)_F`$ is interpreted as a statement that *ascribes a property* to the particle, namely the statement *โThe particle is located in $`S`$.โ* Now take an element $`E(S)๐ช_Q`$. Is it the quantum equivalent of the classical statement $`F^1(S)`$; i.e., an ascription of a property *โlocated in Sโ*? If we interpret $`E(S)`$ this way, then it is trivially true that a quantum particle cannot have a sharp location:
> Let $`\lambda `$. Then $`E(\{\lambda \})=\mathrm{๐}`$ since $`\{\lambda \}`$ has Lebesgue measure zero. Thus, the proposition *โThe particle is located at $`\lambda `$* is a contradiction. *QED*
Thus the property ascription interpretation of elements of $`๐ช_Q`$ rules out our being able to describe particles with sharp locations. However, the property ascription interpretation is not the traditional interpretation of projection operators in quantum theory. According to the traditional interpretation, the projection operators are *observables* or *experimental propositions*. This is sometimes made precise by saying that $`E(S)`$ represents the statement:
> *โA measurement of $`\widehat{Q}`$ will yield a value lying in $`S`$.โ*
Under this interpretation, $`E(\{\lambda \})=\mathrm{๐}`$ does *not* entail that a particle cannot be located at $`\lambda `$; it simply means that it is false (in fact, a contradiction) to say that any measurement of $`\widehat{Q}`$ will yield the precise value $`\lambda `$. (This, for example, may be the result of the fact that precise measurements of $`\widehat{Q}`$ are impossible.) But now, if $`๐ช_Q`$ consists of observables, while $`_F`$ consists of property ascriptions, differences in the state spaces of the two logics do not necessarily reflect differences in the descriptive resources of classical and quantum mechanics.
The distinction between observables (experimental propositions) and property ascriptions should not be drawn only in quantum physics. For example, although $`F^1(\lambda )_F`$ is a classical property ascription, it is doubtful that any classical experiment ever determines that the value of the quantity $`F`$ is the *exact* real number $`\lambda `$; and thus the proposition โ$`\mathrm{Val}(F)=\lambda `$โ is not really an experimental proposition โ even in the classical case. The same intuition is expressed by Birkhoff and von Neumann who claim that, it would be absurd,
> โto call an โexperimental proposition,โ the assertion that the angular momentum (in radians per second) of the earth around the sun was at a particular instant a rational number!โ (p. 825)
Indeed, what experiment would we perform in order to determine that it was a rational number?
There will certainly be a number of different ways to give a mathematically precise account of the notion of approximate measurement. Here, though, I will just explain briefly von Neumannโs \[25, 595-598\] account, without trying to argue positively for its merits. I wish simply to show that according to von Neumannโs account of approximate measurement in *classical* physics, the logic of classical experimental propositions (for one continuous quantity) is mathematically identical to $`๐ช_Q`$.
The fact that measurements are of arbitrarily good, but always finite, precision is given formal statement by von Neumann as follows:
> For each $`n`$, the value space $`\mathrm{\Omega }`$ of $`F`$ may be partitioned into a finite or countably infinite family of measurable sets $`\{N_i^{(n)}:i=1,2,\mathrm{}\}`$, of which each two have measure-zero overlap. We may do this so that the partition $`\{N_i^{(n+1)}\}`$ is strictly finer than $`\{N_i^{(n)}\}`$ for each $`n`$, and if we let $`\delta _n`$ denote the maximum diameter of the sets in $`\{N_i^{(n)}\}`$, then $`\delta _n0`$. We say that some procedure is a $`F`$-measurement of $`n`$-th level precision just in case it determines the $`i`$ such that $`\mathrm{Val}(F)N_i^{(n)}`$. Then, $`F`$-measurements of all finite levels of precision are possible.
With this as the basic principle of measurement, von Neumann proves: For any set $`S`$, we can determine via measurement (up to an arbitrary level of confidence) whether $`\mathrm{Val}(F)S`$, if and only if $`S`$ has non-vanishing Lebesgue measure. Accordingly, it is not possible to determine with any good level of confidence that $`\mathrm{Val}(F)`$ lies in a Lebesgue measure zero set โ in agreement with the intuition that โ$`\mathrm{Val}(F)=\lambda `$โ and *$`\mathrm{Val}(F)`$ is a rational numberโ* are not experimental propositions. Furthermore, if two subsets $`S_1,S_2()`$ agree except possibly on a set of measure zero, then there is no way to determine that $`\mathrm{Val}(F)S_1\backslash S_2`$ or that $`\mathrm{Val}(F)S_2\backslash S_1`$. Thus, two elements $`S_1,S_2_F`$ define the same classical observable just in case $`S_1\mathrm{}S_2=(S_1\backslash S_2)(S_2\backslash S_1)๐ฉ`$. But then the logic of classical observables $`๐ช_F`$ for the quantity $`F`$ is *also* the quotient Boolean algebra $`()/๐ฉ`$. (See p. 825 of Ref. References.)
Now, Teller claims that,
> โโฆif we believe that systems possess exact values for continuous quantities, classical theory contains the descriptive resources for attributing such values to the system, whether or not measurements are taken to be imprecise in some sense. Quantum mechanics has no such descriptive resourceโ (p. 352).
Apparently, then, Teller endorses the claim that classical mechanics does have states assigning exact values to a continuous quantity, even after we take into account the fact that absolutely precise measurements are not possible. However, when we take that into account, the logic of classical experimental propositions $`๐ช_F`$ is *isomorphic* to the logic of quantum experimental propositions $`๐ช_Q`$. Moreover, if we require quantum states to be $`\sigma `$-additive (thereby depriving quantum theory of the resources to describe exact values), we should also require classical states to be $`\sigma `$-additive โ which would also deprive classical theory of the resources to describe exact values. But this conclusion cannot be right: Classical theory does already have these resources to describe exact values, namely points in phase space. The difficulty, then, is not with a lack of resources; but rather with the connection between the theoretical description of point values (viz., pure normal states on $`_F`$) and the โsurfaceโ description given by states on $`๐ช_F`$. I will argue in the next section that the solution to this difficulty is that the โhidden statesโ give rise to finitely-additive (singular) states on $`๐ช_F`$.
### 2.4 Singular states
Let $`๐ฎ`$ be any family of elements in $`()/๐ฉ`$ with the finite meet property. That is, for any finite family $`\{A_1,\mathrm{},A_n\}๐ฎ`$, we have
$$A_1\mathrm{}A_n\mathrm{๐}.$$
(15)
Then, the Ultrafilter Extension Theorem \[19, p. 339\] entails that there is some 2-valued homomorphism $`\omega `$ on $`()/๐ฉ`$ such that $`\omega (A)=1`$ for all $`A๐ฎ`$. In particular, $`()/๐ฉ`$ *does* have pure states, even though they are not $`\sigma `$-additive. Furthermore, any probability distribution $`\rho `$ on $`()/๐ฉ`$ can be written explicitly as a mixture of pure states; i.e., there is some measure $`\mu `$ on the pure state space of $`()/๐ฉ`$ such that
$$\rho =\omega _\alpha ๐\mu (\alpha ).$$
(16)
Thus, the distribution $`\rho `$ does admit an ignorance interpretation; viz., it measures our ignorance of the (singular) state which describes an individual particle in the ensemble.
The pure states on $`()/๐ฉ`$ should be thought of as โsurface states.โ That is, each pure state on $`()/๐ฉ`$ gives a consistent catalog of sure-fire responses of the system to all possible measurements. I will now argue that for each $`\lambda `$, there is at least one surface state whose predictions can be interpreted as consistent with the fact that $`\mathrm{Val}(F)=\lambda `$, and inconsistent with the fact that $`\mathrm{Val}(F)=\xi `$ for any $`\xi \lambda `$.
Consider again the dart throwing example. According to the von Neumann measurement theory, there is no experiment which would tell us the real number $`\lambda `$ at which the dart lands. However, we can rule out segments of the dart board. In particular, choose some open segment $`V`$ along the line. Then, if the system responds $`0`$ to the measurement $`\pi (V)`$, we are entitled to conclude that the dart did not land at any point in the segment $`V`$. In other words, a necessary condition for the value of $`F`$ to be $`\lambda `$ in the state $`\omega `$ is that:
$`\omega (\pi (V))=1,\text{for all}V_\lambda ()`$
where $`_\lambda `$ is the family of open neighborhoods of $`\lambda `$. Suppose, conversely, that $`\omega `$ satisfies $`()`$ with respect to some point $`\lambda `$. \[If $`()`$ holds, we say that the state $`\omega `$ converges to the point $`\lambda `$.\] Then, for any other point $`\xi `$ along the line, there is some open neighborhood $`W`$ of $`\xi `$ which excludes $`\lambda `$. A measurement of $`\pi (W)`$ in the state $`\omega `$ then shows that the dart did not land at $`\xi `$. Thus, when $`\omega `$ converges to $`\lambda `$, the only interpretation consistent with the von Neumann theory of measurement is that $`\mathrm{Val}(F)=\lambda `$; that is, the โhiddenโ state of the system is given by some phase space point along the line $`\{(q,p):F((q,p))=\lambda \}`$.
Now consider again some mixed state $`\rho `$ on $`()/๐ฉ`$. Then, $`\rho `$ is a mixture of pure states as in Eq. 16, and each pure state assigns some definite value to $`F`$. (We do have to be sure that $`\rho `$ falls off sufficiently quickly at infinity. See Sec. 3 of Ref. References.) Thus, this validates the idea that in classical mechanics, the distribution $`\rho `$ represents the โโฆprobability for the value of the measured quantity to have one or another of the exact values in the support of the distribution.โ
However, everything we have just said about singular states of $`()/๐ฉ`$ was entirely neutral as to whether we were discussing classical or quantum mechanics. Thus, a quantum-mechanical probability distribution $`|\psi |^2`$ may also be thought of as representing a probability for $`\widehat{Q}`$ to have one or another of the exact values in the support of $`|\psi |^2`$. According to Teller, though, even if there were some mathematically acceptable method for describing sharp positions in QM, doing so would not be *physically* acceptable:
> โSuch an extended physical theory would describe systems having totally indeterminate momentum, not even highly localized to any finite interval. Such systems would have an infinite expectation value for their kinetic energyโ (p. 353).
This argument, however, is not a criticism of singular states *per se* (which, by the way, are not extensions of the standard formalism). Since the energy observable is an unbounded operator, there are vectors in the Hilbert space that are not in the domain of this observable, and so these vectors too will assign โinfinite kinetic energyโ to the system. Thus, following Tellerโs reasoning here to its logical conclusion, the Hilbert space is not really the state space, rather the domain of the energy observable is the state space. But then, what about other unbounded observables such as the position observable $`\widehat{Q}`$ itself? Should we also throw away all those vectors that assign an infinite expectation value to position? And once weโve cleansed the Hilbert space of vectors not in the domains of all physically relevant observables, would anything be left?
Besides these obvious retorts, it is not completely clear what is meant by Tellerโs talk about assigning infinite expectation values. What we do know โ in the case of pure (singular) states of $`๐ช_Q`$ โ is that these states cannot be thought of as assigning some finite expectation values to the energy observable. (Precisely: Each pure, convergent state on $`๐ช_Q`$ gives rise to a mixed state on the logic of experimental propositions about energy, and measure one of the pure states in its integral decomposition do not converge to a finite value. See Ref. References, Section 3.) It may be mathematically convenient, then, to adjoin $`\mathrm{}`$ to the normal range of values, but we should *not* say that these states assign โ$`\mathrm{}`$โ to the energy observable. Rather, we should think of this infinity โ as we do of the infinities related to singularities in General Relativity โ as indicating the limits of the descriptive capabilities of our theory. And if, with Teller, we take the theory to be descriptively complete, then there is nothing more to describe about energy in these situations.
## 3 Potential difficulties with singular states
In the previous section, I have argued that singular states permit us to give an ignorance interpretation of the probability distributions associated with individual continuous quantities in QM. I imagine my reader may still not be convinced, though, of the acceptability of singular states. In this section, I will confirm that there are indeed difficulties with these singular states. In the final section, however, I will show that these difficulties can be remedied by taking a more modest view of the measurements that can be made on a continuous quantity.
It would be quite natural to think that if we pick a point $`\lambda `$, it determines uniquely a pure state $`\omega `$ on $`()/๐ฉ`$ that converges to $`\lambda `$. There is, however, a difficulty. Suppose we divide the screen (or dartboard) into two halves $`A=\{q:q>\lambda \}`$ and $`B=\{q:q<\lambda \}`$. Then,
$$\pi (A)\pi (B)=\pi (A)\pi (B)\pi (\{\lambda \})=\pi ()=๐,$$
(17)
since $`\{\lambda \}`$ has Lebesgue measure zero. Thus,
$$\omega (\pi (A))+\omega (\pi (B))=1,$$
(18)
while either $`\omega (\pi (A))=0`$ or $`\omega (\pi (B))=0`$. Thus, any pure state $`\omega `$ must assign $`1`$ to either $`\pi (A)`$ or $`\pi (B)`$, but not to both. What would be the right answer for a state in which the particle is located at $`\lambda `$?
Actually, it is easy to show that both value assignments are consistent with the particle being located at $`\lambda `$. Indeed, let $`๐ฎ_1=_\lambda \{\pi (A)\}`$ and let $`๐ฎ_2=_\lambda \{\pi (B)\}`$. Since both $`๐ฎ_1`$ and $`๐ฎ_2`$ have the finite meet property, there is an ultrafilter $`๐ฐ_1`$ which contains $`๐ฎ_1`$ and an ultrafilter $`๐ฐ_2`$ which contains $`๐ฎ_2`$. Since $`๐ฐ_1`$ and $`๐ฐ_2`$ both converge to $`\lambda `$, both should be interpreted as surface manifestations of some โhiddenโ state in which the particle is located at $`\lambda `$. On the other hand, the particleโs being in some hidden state $`(\lambda ,p)`$, gives no information concerning its disposition to respond to a measurement of $`\pi (A)`$.
The apparent mismatch between hidden states and surface states is, in fact, quite severe.
###### Proposition 2.
For each $`\lambda `$, there are at least $`\mathrm{}_0`$ distinct ultrafilters on $`()/๐ฉ`$ that converge to $`\lambda `$.
In order to prove this proposition, we will first require a lemma.
###### Lemma 1.
Suppose that $`\{B_j:jJ\}`$ is a family of open subsets of $``$ such that $`B_iB_j=\mathrm{}`$ when $`ij`$ and such that $`\lambda \mathrm{clo}(B_j)`$ for all $`jJ`$. Then, there is a family $`\{๐ฐ_j:jJ\}`$ of distinct ultrafilters on $`()/๐ฉ`$ such that $`\pi (_\lambda )๐ฐ_j`$ for all $`jJ`$.
###### Proof.
Let $`\{B_j:jJ\}`$ be given as above. For each $`jJ`$, let
$$๐ฎ_j=\pi (_\lambda )\{\pi (B_j)\}.$$
(19)
From the Ultrafilter Extension Theorem, $`๐ฎ_j`$ is contained in a Boolean ultrafilter $`๐ฐ_j`$ if and only if the meet of any finite collection of elements in $`๐ฎ_j`$ is nonzero. Let $`\{\pi (U_1),\pi (U_2),\mathrm{},\pi (U_n)\}๐ฎ_j`$. If $`U_i_\lambda `$ for each $`i`$, then $`_{i=1}^nU_i`$ is an open set that contains $`\lambda `$. If $`U_1=B_j`$, and $`U_i_\lambda `$ for $`i2`$, then $`_{i=2}^nU_i`$ is an open set that contains $`\lambda `$, and since $`\lambda \mathrm{clo}(B_j)`$, $`_{i=1}^nU_i`$ is a nonempty open set. In either case, the open set $`O:=_{i=1}^nU_i`$ is nonempty. Since Lebesgue measure does not vanish on any nonempty open set, we have
$$_{i=1}^n\pi (U_i)=\pi \left(_{i=1}^nU_i\right)=\pi (O)0.$$
(20)
Thus, there is an ultrafilter $`๐ฐ_j`$ that contains $`๐ฎ_j`$. Since $`\pi (B_i)\pi (B_j)=0`$ when $`ij`$, it follows that $`๐ฐ_i๐ฐ_j`$ when $`ij`$. โ
###### Proof of the proposition.
We will consider the case where $`\lambda =0`$. In light of the lemma, it will suffice to construct a family $`\{B_n:n\}`$ of open subsets of $``$ such that $`B_nB_m=\mathrm{}`$ when $`nm`$ and such that $`0\mathrm{clo}(B_n)`$ for all $`n`$.
For each $`n\{0\}`$, let $`a_n=2^n`$. For each $`n,m`$, let $`b_{nm}=(a_n+a_{n1})/2^m`$, and let
$$A_m=\underset{n=1}{\overset{\mathrm{}}{}}(a_n,b_{nm}).$$
(21)
Note that $`A_mA_{m+1}`$ for all $`m`$, and in fact $`A_mA_{m+1}`$ has nonempty interior. Thus, if we set
$$B_m=\mathrm{int}(A_mA_{m+1}),$$
(22)
then $`B_m`$ is a nonempty open set. We must show that $`B_iB_j=\mathrm{}`$ when $`ij`$ and that $`0\mathrm{clo}(B_i)`$ for all $`i`$. Suppose that $`ij`$. We may assume that $`i+1j`$, in which case $`A_{i+1}A_j`$, and
$$B_iB_j\left(A_iA_{i+1}\right)\left(A_jA_{j+1}\right)A_jA_{i+1}=\mathrm{}.$$
(23)
It is easy to verify that the sequence $`\{b_{nm}:n\}`$ is in $`\mathrm{clo}(B_m)`$ for each $`m`$. But $`lim_nb_{nm}=0`$ and since $`\mathrm{clo}(\mathrm{clo}(B_m))=\mathrm{clo}(B_m)`$, it follows that $`0\mathrm{clo}(B_m)`$. โ
Thus, even though we are able to explain probabilistic dispositions as measures of ignorance of sure-fire dispositions, we are not able to further reduce the latter to categorical location properties. So, one could still think that attributing categorical location properties is objectionable on the grounds that these attributions are explanatorily bankrupt.
There is also a serious practical obstacle for using singular states as the explanans for a reduction of probabilistic dispositions; namely, none of these singular states can be explicitly defined. By saying that these singular states cannot be โexplicitly defined,โ I mean to contrast this with the pure normal states of $`()`$ given by principal ultrafilters, as well as with mixed normal states on $`()/๐ฉ`$. First of all, it is quite trivial to give an explicit definition of a principal ultrafilter on $`()`$. Indeed, once I choose some $`\lambda `$, then (trivially) if you give me a Borel set $`S`$, then I can tell you whether or not $`S`$ is in the principal ultrafilter generated by $`\{\lambda \}`$. We have also seen that any quantum-mechanical wavefunction $`\psi `$ defines a mixed normal state $`\rho _\psi `$ on $`()/๐ฉ`$ by means of the formula
$$\rho _\psi (\pi (S))=_S|\psi (q)|^2๐q.$$
(24)
Since it is possible to explicitly define a wavefunction $`\psi `$, Eq. 24 gives an explicit recipe for computing the value $`\rho _\psi (\pi (S))`$ for any Borel set $`S`$.
On the other hand, although we โknowโ that there are ultrafilters (i.e., pure states) on $`()/๐ฉ`$, we do not know this because someone has constructed an example of such an ultrafilter. In order to obtain an ultrafilter, we note that a certain (explicitly defined) family $`๐ฎ`$ of elements in $`()/๐ฉ`$ has the finite meet property, and then we invoke the Ultrafilter Extension Theorem to extend $`๐ฎ`$ to an ultrafilter $`๐ฐ`$. This extension procedure is a classic example of nonconstructive mathematics: We are told that there is some pure state $`\omega `$ on $`()/๐ฉ`$, but we are not given a recipe for determining the value $`\omega (A)`$ for an arbitrary element $`A()/๐ฉ`$.
A strong case can be made that the sometimes imprecise distinction between โgiving an explicit exampleโ and โproving existence (nonconstructively),โ corresponds to a precise distinction in the strength of the set-theoretic axioms used to prove the existence of the object in question (see Ref. References). In particular, many concrete mathematical objects (such as the Real numbers \[17, Appendix\]) are constructed using the Principle of Recursive Constructions:
> (PRC) Suppose $`X`$ is a non-empty set, and let a function $`G`$ be given from the set of finite sequences in $`X`$ into $`๐ซ(X)\backslash \{\mathrm{}\}`$. Then there exists exactly one function $`F:X`$ such that $`F(0)G(\mathrm{})`$ and $`F(n)G(F(0),\mathrm{},F(n1))`$ for all $`n>0`$.
PRC is equivalent in ZF to the axiom of Dependent Choices \[14, p. 147\]:
> (DC) If $`R`$ is a binary relation on a nonempty set $`X`$ such that for every $`xX`$ there exists a $`yX`$ with $`x,yR`$, then there exists a sequence $`(x_n)_n`$ of elements of $`X`$ such that $`x_n,x_{n+1}R`$ for every $`n`$.
DC is a very weak form of the Axiom of Choice: ZF+DC entails the Axiom of Countable Choice, but ZF+DC does not entail the Ultrafilter Extension Theorem, nor does it entail any of the โparadoxicalโ consequences of ZF+AC (e.g. the Banach-Tarski paradox) \[19, Chap. 6\]. In fact, a good case can be made that *applied* mathematics makes use *only* of those objects in the ZF+DC universe . For example, all theorems of classical (19th century) analysis, and even all theorems of the traditional Hilbert space formalism of quantum mechanics can be proved in ZF+DC .
Following , let us say that a mathematical object is *intangible* just in case that object exists in the ZF+AC universe, but that object cannot be proved to exist in the ZF+DC universe. For example, a free ultrafilter on $``$ (i.e., an ultrafilter containing all subsets of $``$ whose complements are finite) is an intangible . We now show that value states of $`()/๐ฉ`$ are also intangibles.
###### Proposition 3.
It cannot be proved in ZF+DC that there is a pure state of $`()/๐ฉ`$.
###### Proof.
Let WUF (weak-ultrafilter principle) denote the statement that there is a free ultrafilter on $``$, and let PS denote the statement that there is a pure state on $`()/๐ฉ`$. We must show that ZF+DC+$`\neg `$PS is consistent. Since ZF+DC+$`\neg `$WUF is consistent , it will suffice to show that ZF+DC$``$(PS$``$WUF). Thus, suppose that PS holds; i.e., there is an ultrafilter $`๐ฐ`$ on $`()/๐ฉ`$. By Prop. 1, $`๐ฐ`$ is not closed under countable meets; i.e., there is a family $`\{P_n:n\}`$ of elements of $`()/๐ฉ`$ such that $`P_n๐ฐ`$ for all $`n`$, but $`_nP_n๐ฐ`$. (Itโs important to note that Prop. 1 uses no choice axiom stronger than DC.) Define a family $`๐ฅ`$ of subsets of $``$ by
$$J๐ฅ_{nJ}P_n๐ฐ.$$
(25)
It is easily verified that $`๐ฅ`$ is an ultrafilter in $`๐ซ()`$ that contains all subsets of $``$ whose complements are finite. โ
## 4 The logic of unsharp experimental <br>propositions
I have argued that it is possible to give an ignorance interpretation of the probabilites associated with a continuous quantity in quantum mechanics. In order to do so, however, I had to make use of singular states, which turn out to have some pretty odd properties that militate against their playing a serious role in explaining a particleโs probabilistic dispositions. Some might attribute the oddities encountered in the last section to the fact that we permitted ourselves to use states that are not $`\sigma `$-additive. On the contrary, I will now argue that the difficulty is not with the states but with the observables: We were trying to account for too many measurement outcomes.
If we take $`()/๐ฉ`$ to be in one-to-one correspondence with possible experiments, then Prop. 2 shows that there are experiments whose outcomes could distinguish two states, both of which correspond to one point $`\lambda `$. If, however, these distinguishing experiments cannot in fact be performed, then these two states corresponding to $`\lambda `$ are empirically equivalent. In this section, I will show how to make this idea precise using the fact that real experimental questions are always โunsharp.โ
Consider again a person, Bob, throwing darts at a one-dimensional dartboard. With the knowledge we have of Bobโs skills, we choose a region $`S`$ on the dartboard which we feel confident Bob can hit on 100 consecutive throws. Since we donโt have time, though, to watch each of Bobโs throws, we rig a device to keep score. In particular, each time a dart hits the board, we get a printout that records $`y`$ when the dart lands in $`S`$ and that records $`n`$ when the dart lands outside of $`S`$. Now, after Bob has made his $`100`$ throws, we take a look at the printout: there are $`98`$ $`y`$โs and $`2`$ $`n`$โs. Do we then conclude that Bob was not as skilled a dart thrower as we thought? Not necessarily. Perhaps it was our score-keeping device, not Bob, which failed on those two counts. Or, perhaps some environmental factor interfered with two of Bobโs throws. In any case, the point is that in a real experimental situation, the measuring device itself and the environment introduce factors of uncertainty which have to be taken into account when interpreting out measurement results.
To make the imperfections in the experimental setup mathematically precise within the confines of the standard Hilbert space formalism of QM, we can associate a confidence measure $`e`$ with the device $``$. (Here I follow Busch et al. .) If, using this measuring device, we attempt to test for the claim โthe particle is in $`S`$,โ then what we actually measure is the โsmearedโ observable
$$E^e(S):=_{}e(q)E(S+q)๐q=(\chi _Se)(\widehat{Q}).$$
(26)
Here the operation $``$ is the convolution product defined by
$$(fg)(q_0):=_{}f(q)g(q_0q)๐q.$$
(27)
For a fixed confidence measure $`e`$, the mapping $`SE^e(S)`$ is a positive operator valued (POV) measure on $`(,())`$. If $`S_1`$ and $`S_2`$ are disjoint Borel sets, then we represent the disjunctive question *โIs the particle in $`S_1S_2`$?โ* formally by
$$E^e(S_1)E^e(S_2):=E^e(S_1)+E^e(S_2).$$
(28)
The conjunctive question *โIs the particle in $`S_1`$ and $`S_2`$?โ* is represented by the operator $`E^e(S_1S_2)`$. In general, however, it is not the case that
$$E^e(S_1S_2)=E^e(S_1)E^e(S_2).$$
(29)
In fact, the operator on the right-hand side of Eq. 29 will not generally be in the range of $`E^e`$. Finally, the question *โIs the particle in the complement of $`S`$?โ* is represented formally by
$$\neg E^e(S):=E^e\left(\backslash S\right).$$
(30)
We obtain the standard projection valued (PV) measure $`E`$ if we apply the above construction to the case in which we have absolute confidence in the accuracy of our measuring apparatus. Formally, if we let $`\delta `$ be the Dirac delta function, then
$$E^\delta (S)=(\chi _S\delta )(\widehat{Q})=\chi _S(\widehat{Q})=E(S),$$
(31)
for all Borel sets $`S`$. On the other hand, if $`e`$ has non-zero deviation (i.e., $`e`$ is an integrable *function*), then for each Borel set $`S`$, $`(\chi _Se)`$ is a uniformly continuous function \[13, Prop. 3.2\].
Realistically, then, the family $``$ of experimental propositions about the location of the particle should consist only of elements of the form $`E^e(S)`$, where $`e`$ is some confidence function with non-zero deviation. Letโs be generous, though, and suppose that $``$ contains all operators of the form $`f(\widehat{Q})`$ where $`f`$ is a uniformly continuous function from $``$ into $`[0,1]`$. If we say that $`AB`$ just in case $`A+B๐`$, then we may extend the exclusive disjunction defined in Eq. 28 by setting
$$AB:=A+B,$$
(32)
when $`AB`$. It is easy to verify then that $`(,,\mathrm{๐},๐)`$ is an *effect algebra*, or *unsharp quantum logic* . (Recall that every Boolean algebra is an effect algebra if we set: $`AB`$ iff. $`AB=\mathrm{๐}`$, and $`AB=AB`$ .) Thus, in particular, for each $`A`$, there is a unique element $`\neg A=๐A`$ such that $`A\neg A`$ and $`A\neg A=๐`$. For $`A,B`$, we say that $`AB`$ just in case there is some $`CA`$ such that $`AC=B`$. A *state* on $``$ is a mapping $`\omega :[0,1]`$ such that $`\omega (๐)=1`$ and
$$\omega (AB)=\omega (A)+\omega (B),$$
(33)
whenever $`AB`$ is defined. From this it follows that whenever $`A,B`$ and $`AB`$, then $`\omega (A)\omega (B)`$ and
$$\omega (BA)=\omega (B)\omega (A).$$
(34)
We say that a state $`\omega `$ of $``$ is pure just in case: If $`\omega =a\rho +(1a)\tau `$, for some $`a(0,1)`$, then $`\omega =\rho =\tau `$. Unlike the Boolean case, however, a pure state of a general effect algebra may take on any value in the interval $`[0,1]`$. (For example, even if Bob hits the bullseye on each try out of $`100`$, our less than ideal score-keeper might only credit him with only, say, $`98`$ hits.)
I began this section with the idea that there are too many observables in $`๐ช_Q()/๐ฉ`$, and that we should be more modest about what can actually be measured. Let me show precisely, now, how the effect algebra $``$ does represent a more modest perspective on what the *experimental* propositions about position truly are.
Both $``$ and $`๐ช_Q`$ are subsets of the algebra $`_Q`$ of all operators of the form $`f(\widehat{Q})`$, where $`f`$ is some Borel function from $``$ into $``$. Although the intersection of $``$ and $`๐ช_Q`$ contains only $`\mathrm{๐}`$ and $`๐`$, we should think of $``$ as containing a much smaller set of observables than $`๐ช_Q`$. Indeed, any element in $``$ may be uniformly approximated by linear combinations of elements in $`๐ช_Q`$ (by the spectral theorem). Thus, any pure state $`\omega `$ on $`()/๐ฉ`$ will give rise to a unique pure state $`\omega |_{}`$ on $``$. (This slight abuse of notation is justified by the fact that pure states on $`๐ช_Q`$ extend uniquely to pure states of the von Neumann algebra $`_Q`$.) On the other hand, since the family of uniformly continous functions is closed under linear combinations and uniform limits, no element of $`()/๐ฉ`$ (other than $`\mathrm{๐}`$ and $`๐`$) can be approximated by linear combinations of elements in $``$. Thus, $``$ is โcoarser grainedโ than $`๐ช_Q`$ in the sense that two states which give different outcomes for measurements in $`๐ช_Q`$ may give identical outcomes for all measurements in $``$. And, indeed, the elements of $``$ do not permit us to distinguish between states that correspond to a common point $`\lambda `$.
###### Proposition 4.
Let $`\omega ,\rho `$ be pure states of $`๐ช_Q`$, both of which converge to $`\lambda `$. Then,
$$\omega (f(\widehat{Q}))=\rho (f(\widehat{Q}))=f(\lambda ),$$
(35)
for all $`f(\widehat{Q})`$.
###### Proof.
See Prop. 3.4 of Ref. References. โ
This proposition gives us everything we want. For each $`\lambda `$, there is precisely one (explicitly defined) pure state $`\omega _\lambda `$ of $``$ corresponding to $`\lambda `$. (Moreover, it can be shown that every โconvergentโ state of $``$ is of this form. See Proposition 5 in the appendix.) Therefore, any statistical state $`\rho _\psi `$ of $``$ decomposes as an integral
$$\rho _\psi =_{}\omega _\lambda ๐\mu (\lambda ),$$
(36)
which permits interpretation as a measure of our ignorance of the precise, categorical location property of individual particles in the ensemble.
Acknowledgments: I would like to thank E. Schechter (Vanderbilt) for helpful correspondence. Special thanks are due to Rob Clifton for providing the impetus for this paper, and for numerous discussions during its composition.
## Appendix A Appendix
###### Lemma 2.
Let $`\omega `$ be a state of $``$. Then $`\omega (aA)=a\omega (A)`$ for all $`A`$ and $`a(0,1)`$.
###### Proof.
Suppose first that $`a=1/2`$. Then, $`\omega (A)\omega ((1/2)A)=\omega ((1/2)A)`$ and so $`\omega ((1/2)A)=(1/2)\omega (A)`$. It then follows easily by induction that $`\omega (aA)=a\omega (A)`$ when $`a=2^n`$ for some $`n`$.
Suppose now that $`a`$ is a diadic rational; i.e., $`a=m/2^n`$ for some $`m,n`$, where $`m<2^n`$. Note then that
$$\frac{m}{2^n}A=\underset{m\text{ times}}{\underset{}{\frac{1}{2^n}A\frac{1}{2^n}A\mathrm{}\frac{1}{2^n}A}},$$
(37)
and so
$$\omega \left(\frac{m}{2^n}A\right)=m\omega \left(\frac{1}{2^n}A\right)=\frac{m}{2^n}\omega (A).$$
(38)
Finally, let $`a`$ be an arbitrary element in $`(0,1)`$. Then, $`a=lim_na_n`$ for some strictly decreasing sequence $`\{a_n\}(0,1)`$ of diadic rationals. Now, for every $`N`$, there is some $`M`$ such that
$$a_nAaA2^N๐,$$
(39)
for all $`nM`$, and therefore
$$\omega \left(a_nAaA\right)2^N,$$
(40)
for all $`nM`$. We may also assume that $`a_na2^N`$ when $`nM`$. Thus,
$`\left|\omega (aA)a\omega (A)\right|`$ $`=`$ $`\left|\omega (aA)\omega (a_nA)+a_n\omega (A)a\omega (A)\right|`$ (41)
$``$ $`\left|\omega (aA)\omega (a_nA)\right|+\left|a_n\omega (A)a\omega (A)\right|`$ (42)
$``$ $`2^N+2^N.`$ (43)
Since this is true for all $`N`$, it follows that $`\omega (aA)=a\omega (A)`$. โ
There are pure states on $``$ that assign $`0`$ to all elements of the form $`E^e(S)`$, where $`S`$ is a compact subset of $``$. (That such states exist can be seen from the proof of Proposition 5 below.) When $`S`$ is compact, then the function $`f:=(\chi _Se)`$ vanishes at infinity; i.e., for each $`ฯต>0`$, there in a $`N`$ such that $`f(x)<ฯต`$ when $`|x|>N`$ \[13, Prop. 3.2\]. This motivates the following definition.
###### Definition.
Let $`\omega `$ be a pure state of $``$. We say that $`\omega `$ converges just in case $`\omega (f(\widehat{Q}))>0`$ for some $`f`$ that vanishes at infinity.
###### Proposition 5.
Let $`\omega `$ be a convergent pure state of $``$. Then, $`\omega (f(\widehat{Q}))=f(\lambda )`$ for some $`\lambda `$.
###### Proof.
Recall that the Stone-Cech compactification $`\beta `$ of $``$ is the unique compact Hausdorff space such that every bounded continuous function $`f:`$ can be extended uniquely to a continuous function $`\overline{f}:\beta `$. Let $`BUC()`$ denote the set of bounded, uniformly continuous functions from $``$ into $``$. Since a uniform limit of elements in $`BUC()`$ is again in $`BUC()`$, it follows that there is a unique compact Hausdorff space $`\gamma `$ such that every $`fBUC()`$ has a unique extension to a continuous function $`\overline{f}:\gamma `$. We may refer to $`\gamma `$ as the uniform compactification of $``$. Let $`C(\gamma )`$ denote the set of continuous functions from $`\gamma `$ into $``$. Thus, there is an algebraic isomorphism from $`BUC()`$ onto $`C(\gamma )`$ and we may identify $``$ with the family of functions in $`C(\gamma )`$ with range $`[0,1]`$.
We show that every (pure) state $`\omega `$ of $``$ has a unique extension to a (pure) state $`\widehat{\omega }`$ of the $`C^{}`$-algebra $`C(\gamma )`$. We may then appeal to the result that any pure state $`\widehat{\omega }`$ on $`C(\gamma )`$ is of the form $`\widehat{\omega }(f)=f(x_0)`$ for some $`x_0\gamma `$ \[16, Corollary 3.4.2\].
Let $`\omega `$ be a state of $``$. Let $`C^+(\gamma )`$ denote the set of functions from $`\gamma `$ into $`[0,+\mathrm{})`$. If $`fC^+(\gamma )`$, then $`f/f`$ and we may define
$$\widehat{\omega }(f):=f\omega \left(f/f\right).$$
(44)
Using Lemma 2, it is easy to see that $`\widehat{\omega }`$ is homogenous with respect to positive real numbers, and it is straightforward to verify that $`\widehat{\omega }`$ is additive over positive functions. Thus, $`\widehat{\omega }`$ is an additive function from $`C^+(\gamma )`$ into $`[0,\mathrm{})`$. It follows then that $`\widehat{\omega }`$ extends uniquely to a linear mapping from $`C(\gamma )`$ into $``$ \[19, Prop. 11.55\].
Suppose now that $`\omega `$ is a pure state of $``$, and let $`\widehat{\omega }`$ be the unique extension of $`\omega `$ to $`C(\gamma )`$ as defined above. Suppose that $`\widehat{\omega }=a\rho +(1a)\tau `$, where $`\rho ,\tau `$ are states of $`C(\gamma )`$ and $`a(0,1)`$. Then $`\rho |_{}`$ and $`\tau |_{}`$ are states of $``$, and $`\omega =a\rho |_{}+(1a)\tau |_{}`$. However, since $`\omega `$ is a pure state of $``$, we have $`\omega =\rho |_{}=\tau |_{}`$ and since extensions are unique, it follows that $`\widehat{\omega }=\rho =\tau `$. Thus, $`\widehat{\omega }`$ is a pure state of $`C(\gamma )`$.
Thus, there is some $`x_0\gamma `$ such that $`\widehat{\omega }(f)=f(x_0)`$ for each $`fC(\gamma )`$. We show that if $`x_0\gamma \backslash `$, then the state $`\omega `$ does not converge. Indeed, since $``$ is dense in $`\gamma `$, there is a net $`(x_a)_{a๐ธ}`$ in $``$ such that $`x_ax_0`$. Let $`f`$ be a function that vanishes at infinity. We show that $`\overline{f}(x_0)=0`$. Let $`ฯต>0`$ be given. Then, there is an $`N`$ such that $`\overline{f}(x)<ฯต`$ for all $`x`$ with $`|x|N`$. However, since the set $`[N,N]`$ is compact in $`\gamma `$ and $`x_0`$, there is some $`b๐ธ`$ such that $`x_a[N,N]`$ for all $`ab`$. Since $`f(x_0)=limf(x_a)`$, $`f(x_0)<ฯต`$. Since this is true for any $`ฯต>0`$, it follows that $`f(x_0)=0`$ and $`\omega (f)=0`$. โ |
warning/0003/hep-th0003283.html | ar5iv | text | # Duffin-Kemmer-Petiau and Klein-Gordon-Fock Equations for Electromagnetic, Yang-Mills and external Gravitational Field Interactions: proof of equivalence.
## 1 Introduction
The question on equivalence of DKP and KGF equations has a long history. In 1926 the second order relativistic equation for spin $`0`$ particle has been discovered by O. Klein , V. Fock , W Gordon and others<sup>1</sup><sup>1</sup>1See where a rich list of references with interesting historical comments can be found, specially about equivalence the DKP and KGF equations. For references see also review , where in particular there is references to works I. Gelfand and A.Yaglom, who obtained the firs order equation for particle with aritrary spin ,too.. After the appearance in 1928 of the famous Dirac equation , G. Petiau , R. Duffin and N. Kemmer proposed in 1936-39 years the first order DKP equation for description of spin $`0`$ and $`1`$ particles. From the beginning of this time to prove equivalence between DKP and KGF equations many attempts have been undertaken (see reference in ). From many questions connected with attempts to proof the equivalence of the both theories we would like to discuss as minimum three important ones: 1. Is there a strict proof of equivalence of the both theories in case of the interaction charged spin $`0`$ particles with eletromagnetic field introduced by minimal way? 2. For which types of interactions is it possible to give strict proof of the equivalence of these both theories? 3. Is there equivalence between these theories on description of process of decay unstable particles (specially decay of $`K_L^0`$ mesons)? For details see references in .
In QED many processes have been investigated in frame work of DKP theories (see works - and others references in ). The main conclusion is that calculation based on DKP and KGF equations yield identical results including one-loop corrections. However the strict proof of equivalence of the both theories for physical matrix elements of $`S`$-matrix does not exist even in QED for any order of the perturbation theory. There was also no attempts to generalize the proof on the other types of interactions. The question about the equivalence of the both theories for description unstable particles is remained open.
The main goal of this paper is to strictly prove that matrix elements of $`S`$-matrix in DKP and KGF theories coincide in cases interaction spin $`0`$ particles with external and quantized Maxwell and Yang-Mills fields and also in case of external gravitational fields (without and with torsion). For the proof we utilized reduction formulas of Lehmann, Symanzik, Zimmermann (LSZ), , and Lagrangian approach to construction the generating functional of GF in all the cases of interactions.
As concerned to description of decay unstable particles, we believe that the both theories have to give the same results as well as for all process which are formulated in terms of renormalizable theories. Although this question goes beyond the scope of the paper we give in conclusion the simple example of Lagranian for description of the decay $`K_L^0`$ meson in DKP theory.
The canonical Hamiltonian approach to quantization of DKP equations has been developed in work where it has been proved that generating functional for GF in DKP theory coincides with that of in Lagrangian approach. Thus we start in this paper from generating functional for GF constracted from Lagrangian action with external sources in exponent of functional integral (named $``$-approach). This approach makes simpler the proof of the equivalence of the both theories .
In section 2 the generating functionals are constructed in DKP theory for GF of spin $`0`$ particles interacted with Maxwell and Yang-Mills fields and with external gravitational field (without torsion and with one). We also prove that all many photons GF (not only matrix elements of $`S`$-matrix) coincide on DKP and KGF theories.
In section 3 the equivalence of the matrix elements of $`S`$-matrix in the both theories are proved for all above mentioned interactions, including non-abelian theories. For strict proof of the results we utilized the wave packets instead of plane waves. In section 4 are given short conclusions about basic results.
## 2 Generating functional
First one apply $``$-approach for construction of the generating functional of GF in DKP theory for charged $`0`$-spin particles interacting with quantized EM field $`A_\mu (x)`$. From general consideration the generating functional in $``$-approach has the following form (in $`\alpha `$-gauge)<sup>2</sup><sup>2</sup>2Notation see in :
$`Z(\overline{๐ฅ},๐ฅ,๐ฅ_\mu )`$ $`=`$ $`Z_0^1{\displaystyle }๐A_\mu ๐\psi ๐\overline{\psi }\mathrm{exp}\{i{\displaystyle }d^4x({\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }{\displaystyle \frac{1}{2\alpha }}\left(_\mu A^\mu \right)^2`$ (1)
$`+๐ฅ_\mu A^\mu +\overline{\psi }(x)(i\beta _\mu ๐^\mu m)\psi (x)+\overline{๐ฅ}\psi +\overline{\psi }๐ฅ)\}`$
where
$$Z_0=Z(0,0,0);D^\mu =^\mu ieA^\mu $$
(2)
After integration over $`\psi \left(x\right)`$ and $`\overline{\psi }\left(x\right)`$ we get:
$`Z(\overline{๐ฅ},๐ฅ,๐ฅ_\mu )`$ $`=`$ $`Z_0^1{\displaystyle }๐A_\mu \mathrm{exp}\{i{\displaystyle }d^4x({\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }{\displaystyle \frac{1}{2\alpha }}\left(_\mu A^\mu \right)^2`$ (3)
$`+Tr\mathrm{ln}S(x,x,A){\displaystyle }d^4y\overline{๐ฅ}\left(x\right)S(x,y,A)๐ฅ\left(y\right))\}`$
here
$$S(x,y,A)=\left(i\beta _\mu D^\mu m\right)^1\delta ^4\left(xy\right)$$
(4)
is GF of particles in the field $`A_\mu \left(x\right)`$; the term $`Tr\mathrm{ln}S(x,x,A)`$ is responsible for arising of the all diagrams of vacuum polarization.This term can be transformed to the following view<sup>3</sup><sup>3</sup>3We utilized the equation (24), which expresses $`\psi _\alpha (x),\overline{\psi }_\alpha (x)`$ through the components.,
$`detS(x,y,A)`$ $`=`$ $`C{\displaystyle ๐\psi ๐\overline{\psi }\mathrm{exp}\left\{i\overline{\psi }\left(i\beta _\mu D^\mu m\right)\psi \right\}}`$
$`=`$ $`C{\displaystyle }\underset{๐ผ}{\mathrm{\Pi }}๐\phi _\alpha ๐\phi _\alpha ^{}\mathrm{exp}\{i{\displaystyle }d^4x(\phi ^{}D_\mu \phi ^\mu +\phi ^\mu D_\mu \phi `$
$`m(\phi ^{}\phi +\phi ^\mu \phi _\mu ))\}`$
$`=`$ $`C{\displaystyle ๐\phi ^{}๐\phi \mathrm{exp}\left\{\frac{i}{m}d^4x\phi ^{}\left(D_\mu D^\mu +m^2\right)\phi \right\}}`$
$`=`$ $`detG(x,y,A)=\mathrm{exp}Tr\mathrm{ln}G(x,x,A)`$
where
$$G(x,y,A)=\left(D_\mu D^\mu +m^2\right)^1\delta ^4\left(xy\right)$$
(6)
is the GF of KGF equation in EM field $`A_\mu (x)`$. Thus utilizing Eqs. (3), (2), (6) we get
$`Z(\overline{๐ฅ},๐ฅ,๐ฅ_\mu )`$ $`=`$ $`Z_0^1{\displaystyle }๐A_\mu \mathrm{exp}\{i{\displaystyle }d^4x({\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }{\displaystyle \frac{1}{2\alpha }}\left(_\mu A^\mu \right)^2`$
$`+๐ฅ_\mu A^\mu +Tr\mathrm{ln}G(x,x,A){\displaystyle }d^4y\overline{๐ฅ}\left(x\right)S(x,y,A)๐ฅ\left(y\right))\}`$
From Eqs. (6), (2) we get to important conclusions: The all many photons GF (Not only matrix elements of $`S`$-matrix for real photons ) coincide in DKP and KGF theories.
In general case one can obtain the generating functional or GF in external $`A_\mu ^{ex}`$ by sustitution in Eqs. (1), (2)
$$D^\mu ^\mu ie\left(A^\mu +A_{ex}^\mu \right)$$
(8)
However, if $`A_{ex}^\mu `$ is not so strong to create pairs of particles, the generating functional in $`A_{ex}^\mu `$ will be equal to:
$$Z^{ex}=\mathrm{exp}\left\{id^4xd^4y\overline{๐ฅ}\left(x\right)S(x,y,A^{ex})๐ฅ\left(y\right)\right\}$$
(9)
and particles do not interact betweem themselves but only with $`A_{ex}^\mu `$.
Now we consider interaction DKP spin-$`0`$ charged particles with external gravitational fields. The Lagrangian density in this case can be written in the form
$``$ $`=`$ $`\sqrt{g}\{{\displaystyle \frac{i}{2}}[\overline{\psi }\beta ^\mu (_\mu +{\displaystyle \frac{1}{2}}\omega _\mu ^{ab}S_{ab})\psi `$
$`(_\mu \overline{\psi }{\displaystyle \frac{1}{2}}\omega _\mu ^{ab}\overline{\psi }S_{ab})\beta ^\mu \psi m\overline{\psi }\psi ]\}`$
Here as in case Dirac equation, we have to introduce nontrivial tetrad field $`e_a^\mu \left(x\right)`$, which can be used to define the Riemannian metric ($`a`$-vector Lorentz index; $`\mu `$-riemannian one):
$$g_{\mu \nu }=e_\mu ^ae_\nu ^b\eta _{ab},\eta _{ab}=diag\{1,1,1,1\}$$
(11)
with properties:
$`e_\mu ^ae_a^\nu `$ $`=`$ $`\delta _\mu ^\nu ,e_\mu ^ae^{\mu b}=\eta ^{ab}`$ (12)
$`\sqrt{g}`$ $`=`$ $`dete_a^\mu =e`$
$`\beta _\mu =e_\mu ^a\beta _a`$, where $`5\times 5`$ matrix $`\beta _a`$ satisfy usual DKP commutation relations ; the connection $`\omega _\mu ^{ab}`$ equal to (see for instance the book ):
$$\omega _\mu ^{ab}=e_\lambda ^ae^{\nu b}\mathrm{\Gamma }_{\mu \nu }^\lambda e^{\nu b}_\mu e_\nu ^a$$
(13)
We consider two different cases: without torsion or with one. Without torsion Levi-Civita connection is
$$\mathrm{\Gamma }_{\mu \nu }^\lambda =\mathrm{\Gamma }_{\nu \mu }^\lambda =\stackrel{}{\mathrm{\Gamma }}_{\mu \nu }^\lambda =\frac{1}{2}\stackrel{}{g}^{\lambda \rho }\left(_\mu \stackrel{}{g}_{\rho \nu }+_\nu \stackrel{}{g}_{\rho \mu }_\rho \stackrel{}{g}_{\mu \nu }\right)$$
(14)
and
$$_\mu \sqrt{g}=\frac{\sqrt{g}}{2}g^{\rho \lambda }_\mu g_{\rho \lambda }\sqrt{g}\stackrel{๐}{\mathrm{\Gamma }}_{\rho \mu }^\rho $$
(15)
If torsion is not equal zero $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ can be defined as Cartan connection, :
$$\mathrm{\Gamma }_{\mu \nu }^\lambda =e_a^\lambda _\nu e_\mu ^a$$
(16)
Now from Eqs. (13) and (16) we obtain:
$$\omega _\mu ^{ab}=e_\lambda ^be^{\nu a}\left(\mathrm{\Gamma }_{\mu \nu }^\lambda \mathrm{\Gamma }_{\nu \mu }^\lambda \right)$$
(17)
or
$$\omega _{\mu \nu }^\lambda =e_{\nu a}e_b^\lambda \omega _\mu ^{ab}=\mathrm{\Gamma }_{\mu \nu }^\lambda \mathrm{\Gamma }_{\nu \mu }^\lambda 2Q_{\mu \rho }^\lambda $$
(18)
where the torsion $`Q_{\mu \rho }^\lambda `$ is a third rank tensor . Thus we have two different Lagrangian depending on Eqs. (14) or (16) which define the $`\stackrel{๐}{\omega }_\mu ^{ab}`$ or $`\omega _\mu ^{ab}`$ in Eq. (17). However, in both cases from Eq. (2) we get the form-equivalent equations for $`\psi `$ (The equation for $`\psi \left(x\right)`$ in case Einstein-Cartan gravity is discussed shortly in p.2 in Conclusion):
$$\left(i\beta ^\mu \left(_\mu +\frac{1}{2}\omega _\mu ^{ab}S_{ab}\right)m\right)\psi \left(i\beta _\mu ^\mu m\right)\psi =0$$
(19)
To show it in case (17) when for instance torsion not equals zero we can transform Eq. (2) (omiting total derivatives) to the following view:
$``$ $`=`$ $`\sqrt{g}\{i\overline{\psi }(\beta _\mu ^\mu +{\displaystyle \frac{1}{2}}\omega _\mu ^{ab}\beta ^\mu S_{ab}m)\psi `$
$`+{\displaystyle \frac{i}{2}}\overline{\psi }(_\mu \left(\beta ^\mu \sqrt{g}\right)\omega _\mu ^{\mu b}\beta _b)\psi \}`$
We utilize at the derivation of Eq. (2) the following equation:
$$\omega _\mu ^{ab}S_{ab}\beta ^\mu =\omega _\mu ^{ab}\beta ^\mu S_{ab}2\omega _\mu ^{\mu a}\beta _a$$
(21)
One shows that the last term in Eq. (2) equals zero:
$`_\mu \left(\sqrt{g}\beta ^\mu \right)`$ $`=`$ $`\beta ^a_\mu \left(e_a^\mu e\right)=\beta ^a\left[\left(_\mu e_a^\mu \right)e+e_a^\mu _\mu e\right]`$
$`=`$ $`e\beta ^a\left[\left(_\mu e_a^\mu \right)+e^\mu \mathrm{\Gamma }_{\rho \mu }^\rho \right]=e\beta ^a\omega _{\mu a}^be_b^\mu =e\omega _\mu ^{\mu b}\beta _b`$
where we used that $`\left(_\mu +\omega _{\mu a}^b\mathrm{\Gamma }_{\lambda \mu }^\lambda \right)e_a^\mu =0`$. Thus the last term in Eq. (2) equals zero. We get the same answer without torsion, too. Then Lagrangian (2) can be written so:
$$=\sqrt{g}\left\{i\overline{\psi }\left(\beta _\mu ^\mu m\right)\psi \right\}$$
(23)
One writes down the Lagrangian (2),(23) in component form utilizing the expressions for $`\psi _\alpha (x),\overline{\psi }_\alpha (x)`$:
$`\psi _\alpha `$ $`=`$ $`(\phi ,\phi ^0,\phi ^1,\phi ^2,\phi ^3,)`$ (24)
$`\overline{\psi }_\alpha `$ $`=`$ $`\left(\psi ^{}\eta \right)_\alpha =(\phi ^{},\phi ^0,\phi ^1,\phi ^2,\phi ^3,)`$
we obtain<sup>4</sup><sup>4</sup>4For neutral particle $`=\frac{1}{2}\sqrt{g}\left\{\phi D_\mu \phi ^\mu +\phi ^\mu _\mu \phi m^2\left(\phi ^2+\phi _\mu \phi ^\mu \right)\right\}`$
$$L=\sqrt{g}\left\{\phi ^{}D_\mu \phi ^\mu +\phi ^\mu _\mu \phi m\left(\phi ^2+\phi _\mu ^{}\phi ^\mu \right)\right\}$$
(25)
where
$$D_\mu =_\mu +\mathrm{\Gamma }_{\nu \mu }^\nu $$
(26)
From Eqs. (25), (26) we get the following $``$-equations :
$`D_\mu \phi ^\mu +m\phi `$ $`=`$ $`0,_\mu \phi +m\phi _\mu =0`$ (27)
$`D_\mu ^\mu \phi +m^2\phi `$ $`=`$ $`0`$ (28)
The same equations arise for $`\phi ^{}`$.
If $`D_\mu =\stackrel{}{D}_\mu _\mu +\stackrel{}{\mathrm{\Gamma }}_{\nu \mu }^\nu `$ one gets KGF equation without torsion (see Eqs. (18), (19) in ); if $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ is defined by Eq. (16) we get equation for scalar particles with torsion (see (27), (28) in ).
Thus on the classical level it is proved equivalence of DKP and KGF equations in external gravitational fields .
On the quantum level ($`2^{nd}`$-quantization) we must construct (also as in external $`A^{ex}`$) generating functional for GF of spin-$`0`$ particles, interacting with external gravitational field $`e_\mu ^a\left(x\right)`$. By definition the such generating functional in $``$-approach is (see Eq. (23)):
$$Z(\overline{๐ฅ},๐ฅ)=Z_0^1๐\psi ๐\overline{\psi }\mathrm{exp}\left\{id^4x\sqrt{g}\left(\overline{\psi }\left(i\beta _\mu ^\mu m\right)\psi +\overline{๐ฅ}\psi +\overline{\psi }๐ฅ\right)\right\}$$
(29)
where $`Z_0=Z(0,0),`$ $`\overline{๐ฅ}`$, $`๐ฅ`$ are external currents.
Integrating in Eq. (29) over $`\overline{\psi }`$ and $`\psi `$ we obtain:
$$Z(\overline{๐ฅ},๐ฅ)=\mathrm{exp}\left\{id^4xd^4ye\left(x\right)e\left(y\right)\overline{๐ฅ}(x)S(x,y,e)๐ฅ\left(y\right)\right\}$$
(30)
Here we have introduced the total GF of DKP particle in external gravitational field $`e_a^\mu `$:
$$S(x,y,e)=\left(i\beta _\mu _x^\mu m\right)^1\delta ^4\left(xy\right)e^1\left(y\right)$$
(31)
From Eq. (29) we also have:
$$S(x,y,e)=i0\left|T\widehat{\psi }\left(x\right)\widehat{\overline{\psi }}\left(y\right)\right|0$$
(32)
We suppose that external $`e_\mu ^a\left(x\right)`$ is enough weak to create pairs of new particles. Thus one can consider, in this case, only one particle GF in external $`e_\mu ^a\left(x\right)`$, since each particle interacts with $`e_\mu ^a\left(x\right)`$, but not between themselves.
## 3 Equivalence between physical matrix elements of S-matrix in DKP and KGF equations
We use LSZ reduction formulas for proof of the equivalence of the both theories. The main goal of LSZ approach is to express the matrix elements of $`S`$-matrix through total many particles GF, for instance, Eq. (2), making minimal assumptions as possible. To apply this approach to DKP theory we must to write down the operatoreโs solution of the free DKP equations. Taking into account that in DKP theory there are only two linearly independent solutions of the free equation ,one can write:
$`\widehat{\psi }_{\begin{array}{c}in\\ out\end{array}}\left(x\right)`$ $`=`$ $`{\displaystyle \frac{1}{\left(2\pi \right)^{3/2}}}{\displaystyle d^3p\left\{u^{}\left(๐ฉ\right)\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)e^{ipx}+u^+\left(๐ฉ\right)\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right)e^{ipx}\right\}}`$ (33)
$`\widehat{\overline{\psi }}_{\begin{array}{c}in\\ out\end{array}}\left(x\right)`$ $`=`$ $`{\displaystyle \frac{1}{\left(2\pi \right)^{3/2}}}{\displaystyle d^3p\left\{\overline{u}^{}\left(๐ฉ\right)\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right)e^{ipx}+\overline{u}^+\left(๐ฉ\right)\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)e^{ipx}\right\}}`$ (34)
here
$`\left(\widehat{p}m\right)u^\pm \left(๐ฉ\right)`$ $`=`$ $`0,\overline{u}^{}\left(๐ฉ\right)\left(\widehat{p}\pm m\right)=0,\widehat{p}=\beta _\mu p^\mu `$ (35)
$`p_0`$ $`=`$ $`\left(๐ฉ^2+m^2\right)^{1/2}\omega \left(๐ฉ\right)=\omega `$
operators $`\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^\pm ,`$ $`\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^\pm `$ satisfly to usual commutation relations and the solutions in component from are:
$$u_\alpha ^\pm =\sqrt{\frac{m}{2\omega }}(1,\pm \frac{i\omega }{m},\pm \frac{ip^1}{m},\pm \frac{ip^2}{m},\pm \frac{ip^3}{m})$$
(36)
It is easy to check the scalar products are:
$`u^{}(๐ฉ)\beta _0u^{}(๐ฉ)`$ $`=`$ $`\overline{u}^{}(๐ฉ)\beta _0u^{}(๐ฉ)=\pm 1`$ (37)
$`\overline{u}^\pm (๐ฉ)\beta _0u^{}(๐ฉ)`$ $`=`$ $`0`$
Operatores $`\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^\pm ,`$ $`\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^\pm `$ are that of creation and annihilation with positive and negative charges accordingly.
From Eqs. (33) and (34) one also has:
$`\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)`$ $`=`$ $`{\displaystyle \frac{1}{\left(2\pi \right)^{3/2}}}{\displaystyle d^3xe^{ipx}\overline{u}^{}(๐ฉ)\beta _0\widehat{\psi }_{\begin{array}{c}in\\ out\end{array}}\left(x\right)}`$ (38)
$`\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right)`$ $`=`$ $`{\displaystyle \frac{1}{\left(2\pi \right)^{3/2}}}{\displaystyle d^3xe^{ipx}\overline{u}^+(๐ฉ)\beta _0\widehat{\psi }_{\begin{array}{c}in\\ out\end{array}}\left(x\right)}`$ (39)
$`\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right)`$ $`=`$ $`\left(\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)\right)^{},\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)=\left(\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right)\right)^{}`$ (40)
However, decomposition (33), (34) of the operators $`\widehat{\psi }_{\begin{array}{c}in\\ out\end{array}}\left(x\right)`$, $`\widehat{\overline{\psi }}_{\begin{array}{c}in\\ out\end{array}}\left(x\right)`$ over the operators $`\widehat{a}_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ฉ\right)`$, $`\widehat{b}_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ฉ\right)`$, which are that of creation and annihilation of the plane wave do not give possibility to construct the total set of the normalizale physical states in Hilbert space. Moreover, utilizing these states for construction reduction formulas of LSZ we can not carry out strictly matematically proof of the equivalence of DKP and KGF theories. For these aims we have to construct new operators which create and annihilate the wave packets states. We define the new operators in the following decomposition (we omit sign ^ over operators)
$`a_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}f_n\left(๐ฉ\right)a_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ง\right)`$ (41)
$`b_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}f_n\left(๐ฉ\right)b_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ง\right)`$
Here $`f_n\left(๐ฉ\right)=`$ $`f_n^{}\left(๐ฉ\right)`$ are total ortonormalized set of functions with properties<sup>5</sup><sup>5</sup>5We choose $`f_n\left(๐ฉ\right)=f_n^{}\left(๐ฉ\right)=\mathrm{exp}\left(\frac{๐ฉ^2}{2m^2}\right)\mathrm{\Pi }_{i=1}^3H_n\left(\frac{p_i}{m}\right)`$, where $`H_n\left(x\right)`$ denotes a Hermite polinomial of degree $`n`$.
$$f_n\left(๐ฉ\right)f_m\left(๐ฉ\right)d^3๐ฉ=\delta _{nm},\underset{n=0}{\overset{\mathrm{}}{}}f_n\left(๐ฉ\right)f_n\left(๐ช\right)=\delta ^3\left(๐ฉ๐ช\right)$$
(42)
from Eqs.(41) and (42) we get:
$$a_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ง\right)=d^3๐ฉf_n\left(๐ฉ\right)a_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ฉ\right),b_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ง\right)=d^3๐ฉf_n\left(๐ฉ\right)b_{\begin{array}{c}in\\ out\end{array}}^\pm \left(๐ฉ\right)$$
(43)
From commutations relations:
$$[a_{\begin{array}{c}in\\ out\end{array}}^{}\left(๐ฉ\right),a_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ช\right)]=[b_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ฉ\right),b_{\begin{array}{c}in\\ out\end{array}}^+\left(๐ช\right)]=\delta ^3\left(๐ฉ๐ช\right)$$
(44)
and from Eqs. (41)-(43) one get:
$$[a_{\begin{array}{c}in\\ out\end{array}}^{}\left(n\right),a_{\begin{array}{c}in\\ out\end{array}}^+\left(m\right)]=[b_{\begin{array}{c}in\\ out\end{array}}^{}\left(n\right),b_{\begin{array}{c}in\\ out\end{array}}^+\left(m\right)]=\delta _{mn}$$
(45)
Vaccuum and $`S`$-matrix are defined as usual
$`a_{\begin{array}{c}in\\ out\end{array}}^{}\left(n\right)|0_{\begin{array}{c}in\\ out\end{array}}`$ $`=`$ $`b_{\begin{array}{c}in\\ out\end{array}}^{}\left(n\right)|0_{\begin{array}{c}in\\ out\end{array}}=0,S|0_{\begin{array}{c}in\\ out\end{array}}=|0_{\begin{array}{c}in\\ out\end{array}}`$ (46)
$`a_{out}^+`$ $`=`$ $`a_{in}^+S`$ (47)
The any physical matrix element of $`S`$-matrix one has:
$$n^{},m^{};out|n,m;in=n^{},m^{};in\left|S\right|n,m;in$$
(48)
were $`n+m=n^{}+m^{}`$ is conservation of charge and
$$|n,m;_{out}^{in}\underset{i=1}{\overset{๐}{\mathrm{\Pi }}}\underset{j=1}{\overset{๐}{\mathrm{\Pi }}}a_{\begin{array}{c}in\\ out\end{array}}^+\left(n_i\right)b_{\begin{array}{c}in\\ out\end{array}}^+\left(n_i\right)|0$$
(49)
Now we can formulate main assumption of LSZ approach<sup>6</sup><sup>6</sup>6This assumption can be proved in case of microcasuality theories, when comutator $`[\psi \left(x\right),\overline{\psi }\left(y\right)]=0,\left(xy\right)^2<0`$: for any matrix elemnts of Heisenberg operators $`\widehat{\psi }\left(x\right)`$ and $`\widehat{\overline{\psi }}\left(x\right)`$ the following asymptoptic relation are implemented:
$$\underset{x_0=\mathrm{}}{lim}n^{},m^{};_{out}^{in}\left|\widehat{\psi }\left(x\right)\right|n,m;_{out}^{in}=n^{},m^{};_{out}^{in}\left|\widehat{\psi }_{\begin{array}{c}in\\ out\end{array}}\left(x\right)\right|n,m;_{out}^{in}$$
(50)
and the same relations for $`\widehat{\overline{\psi }}\left(x\right)`$.
In order to do not complicate the proof of the equivalence of the DKP and KGF theories we restricted by considering of the matrix elements of $`S`$-matrix for particles with the same (positive) charges and one utilizes LSZ reduction formula. We have:
$`0\left|\underset{i=1}{\overset{๐}{\mathrm{\Pi }}}a_{out}^{}\left(n_i\right)\underset{j=1}{\overset{๐}{\mathrm{\Pi }}}a_{in}^+\left(m_j\right)\right|0`$ $`=`$ $`0\left|\underset{i=1}{\overset{k1}{\mathrm{\Pi }}}a_{out}^{}\left(n_i\right)a_{out}^{}\left(n_k\right)\underset{j=1}{\overset{๐}{\mathrm{\Pi }}}a_{in}^+\left(m_j\right)\right|0`$ (51)
$`=`$ $`Cn_1,\mathrm{},n_{k1};out|{\displaystyle d^4xf_{n_k}^{}\left(x\right)}`$
$`\left(i\beta _\mu \stackrel{}{}_x^\mu m\right)\widehat{\psi }\left(x\right)|n_1,\mathrm{},n_k;in`$
Here
$$f_{\alpha ;n_k}^{}\left(x\right)=\frac{1}{\left(2\pi \right)^{3/2}}d^3p\overline{u}_\alpha ^{}\left(๐ฉ\right)f_{n_k}\left(๐ฉ\right)e^{ipx},p_0=\omega \left(๐ฉ\right)$$
(52)
and $`C`$\- not essential constant
Utilizing Eqs. (24), (36), (52) we can rewrite Eq. (51) in component form:
$`n_1,\mathrm{},n_{k1};out|{\displaystyle }d^4xd^3๐ฉf_{n_k}\left(๐ฉ\right)\{{\displaystyle \frac{1}{m}}e^{ipx}(\stackrel{}{\mathrm{}}_x+m^2)\widehat{\phi }\left(x\right)`$ (53)
$`+{\displaystyle \frac{}{x^\mu }}\left[e^{ipx}({\displaystyle \frac{1}{m}}_x^\mu \widehat{\phi }\left(x\right)\widehat{\phi }^\mu \left(x\right))\right]\}|n_1,\mathrm{},n_k;in`$
The main idea of the proof is to show that the second term under total derivative in Eq. (53) is equal to zero. We have<sup>7</sup><sup>7</sup>7Itโs easy to show that term in (3) under total derivative has no $`\delta `$-function singularities.
$`{\displaystyle d^4xd^3๐ฉf_{n_k}\left(๐ฉ\right)\frac{}{x^\mu }\left[e^{ipx}\left(\frac{1}{m}_x^\mu \widehat{\phi }\left(x\right)\widehat{\phi }^\mu \left(x\right)\right)\right]}`$
$`=`$ $`{\displaystyle ๐\sigma _\mu d^3๐ฉf_{n_k}\left(๐ฉ\right)e^{ipx}\left(\frac{1}{m}_x^\mu \widehat{\phi }\left(x\right)\widehat{\phi }^\mu \left(x\right)\right)}`$
One chooses the surface $`\sigma _\mu `$ so:
$$\sigma _\mu :\{Tx_0T;Lx_iL;i=1,2,3\}$$
(55)
The firs term, $`\mu =0`$ equals
$`{\displaystyle }d^3x[{\displaystyle }d^3๐ฉf_{n_k}\left(๐ฉ\right)e^{i\omega T+i\mathrm{๐ฉ๐ฑ}}({\displaystyle \frac{1}{m}}{\displaystyle \frac{}{T}}\widehat{\phi }(๐ฑ,T)\widehat{\phi }^0(๐ฑ,T))`$ (56)
$`{\displaystyle }d^3๐ฉf_{n_k}\left(๐ฉ\right)e^{i\omega T+i\mathrm{๐ฉ๐ฑ}}({\displaystyle \frac{1}{m}}{\displaystyle \frac{}{T}}\widehat{\phi }(๐ฑ,T)\widehat{\phi }^0(๐ฑ,T))]`$
Since
$$\underset{T\mathrm{}}{lim}\frac{}{T}\widehat{\phi }(๐ฑ,\pm T)=\frac{}{T}\widehat{\phi }_{\begin{array}{c}in\\ out\end{array}}(๐ฑ,\pm T)=\widehat{\phi }_{\begin{array}{c}in\\ out\end{array}}^0(๐ฑ,\pm T)$$
(57)
the first term (56) disappears in limit $`T\pm \mathrm{}`$.
It is enough to consider only one term in Eqs. (3) and (55), for instance $`i=1`$, and to show that it goes to zero at $`L\pm \mathrm{}`$ due to Gaussian properties of the packets $`f_n\left(๐ฉ\right)`$ (see<sup>5</sup> p.8)
The term $`\mu =1`$ is:
$`I(L)={\displaystyle }dx_0dx_{}{\displaystyle }d^3๐ฉf\left(๐ฉ\right)e^{i\omega x_0ip_{}x_{}}\{e^{ip_1x_1}({\displaystyle \frac{1}{m}}_x^1\widehat{\phi }\left(x\right)\widehat{\phi }^1\left(x\right))_{x_1=L}`$
$`e^{ip_1x_1}({\displaystyle \frac{1}{m}}_x^1\widehat{\phi }\left(x\right)\widehat{\phi }^1\left(x\right))_{x_1=L}\}`$ (58)
where $`p_i=(p_1,p_{}),`$ $`x_i=(x_1,x_{})`$, $`f_n=f_{n_k}`$
One estimates the first term $`\left(x_1=L\right)`$ in Eq. (58), which can be writen in the form:
$`I\left(L\right)`$ $``$ $`{\displaystyle ๐p_1e^{ip_1L}e^{\frac{p_1^2}{2m^2}}H_n\left(\frac{p_1}{m}\right)๐x_0๐x_{}๐๐ฉ_{}f_n\left(๐ฉ_{}\right)e^{i\omega x_0ip_{}x_{}}}`$ (59)
$`\left({\displaystyle \frac{1}{m}}_x^1\widehat{\phi }\left(x\right)\widehat{\phi }^1\left(x\right)\right)`$
By definition the integral over $`x_0,x_{},๐ฉ_{}`$ in Eq. (59) is a distribution (generalized function) of polinomial grows or Shwartz distribution, (renormalizable theory), i.e. it can grow not faster than polinomial: $`C_1p_1^l|L|^k`$, where $`l`$ and $`k`$ are some numbers, $`C_1`$ is a constant.
Then
$`I\left(L\right)`$ $``$ $`C_1|L|^k{\displaystyle ๐p_1p_1^le^{ip_1L}e^{\frac{p_1^2}{2m^2}}H_n\left(\frac{p_1}{m}\right)}`$
$`<`$ $`C_1|L|^k\left|\left(i{\displaystyle \frac{}{L}}\right)^{l+n}\right|{\displaystyle ๐p_1e^{ip_1L\frac{p_1^2}{2m^2}}}`$
$`=`$ $`C_1|L|^k\left|\left(i{\displaystyle \frac{}{L}}\right)^{l+n}\right|e^{\left(Lm\right)^2}`$
Thus $`I\left(L\right)`$ decreases as $`e^{\left(Lm\right)^2}|L|^{k+l+n}`$, and we proved that the second term in Eq. (53) equals zero.
If we repeat the LSZ procedure for the second operator $`a_{out}^+\left(n\right)`$ in Eq. (51) we get
$`k;out|k;in`$ $`=`$ $`Ck2;out|{\displaystyle d^4x_1d^4x_2d^3๐ฉ_1d^3๐ฉ_2e^{ip_1x_1}f_{n_1}^{}\left(๐ฉ_1\right)f_{n_2}\left(๐ฉ_2\right)}`$
$`\left({\displaystyle \frac{\stackrel{}{\mathrm{}}_1+m^2}{m}}\right)\{e^{ip_2x_2}\left({\displaystyle \frac{\stackrel{}{\mathrm{}}_2+m^2}{m}}\right)T\widehat{\phi }\left(x_1\right)\widehat{\phi }\left(x_2\right)`$
$`+_\mu ^{x_2}\left[e^{ip_2x_2}(_{x_2}^\mu T\widehat{\phi }\left(x_1\right)\widehat{\phi }\left(x_2\right)T\widehat{\phi }\left(x_1\right)\widehat{\phi }\left(x_2\right))\right]\}|k;in`$
Agian one can prove the last term under total derivative equals zero due to properties of packets $`f_n(๐ช)`$. Continuating this inductive procedure we go to conclusion that all physical matrix elements of $`S`$-matrix in DKP theory coincide with that of in KGF theory independently of character interaction on these theories, if in the both theories the LSZ asymptotic conditions (50) are implemented for Heisenberg operators<sup>8</sup><sup>8</sup>8The quasilocal terms do not appear in Eq. (LABEL:ec63) ..
From general result (LABEL:ec63) the equivalence between DKP and KGF theories also follows in case interaction spin-$`0`$ particles with external eletromagnetic field, $`A_\mu ^{ex}\left(x\right)`$, and gravitational fields, $`e_{\mu a}^{ex}\left(x\right)`$, see Eqs. (9) and (30)-(32).
Now we briefly describe the proof of equivalence of the theories in case interaction spin-$`0`$ particles with non-abelian (Yang-Mills) external and quantized fields $`A_\mu ^i`$, where $`i=1,\mathrm{},2N^21`$, is the group index of $`SU(N)`$ group.
One is restricted by case when spin-o particles in DKP and KGF theories are taken on fundamental representation of the $`SU(N)`$ group.
The initial density of Lagrangian in DKP theory is (in $`\alpha `$-gauge)
$`_{DKP}`$ $`=`$ $`\overline{\psi }_\alpha \left(x\right)\left(i\beta _\mu D_{ab}^\mu m\delta _{ab}\right)\psi _b{\displaystyle \frac{1}{4}}F_{\mu \nu }^iF_i^{\mu \nu }{\displaystyle \frac{1}{2\alpha }}\left(_\mu A_i^\mu \right)^2`$
$`\overline{C}_a\left(_\mu D_{ab}^\mu C_b\right)`$
Here: $`\left(D^\mu \right)_{ab}=^\mu \delta _{ab}ig\left(A_i^\mu T_i\right)_{ab},`$ $`a=1,\mathrm{}N`$, index of fundamental representation. $`F_{\mu \nu }^i=_\mu A_\nu ^i_\nu A_\mu ^i+igf^{ijk}A_\mu ^jA_\nu ^k;`$ $`A_\mu ^i`$-is Yang-Mills field; $`T_i`$ are $`N\times N`$ matrixes with the commutation relation
$$[T_i,T_j]=if_{ijk}T_k$$
(63)
$`f_{ijk}`$ are structure constant of $`SU(N)`$ group; $`\overline{C}_a,C_b`$ Faddeev-Popov anticommuting ghost fields. We also write down only spin-$`0`$ part of $`_{int}`$ in KGF theory:
$$_{KGF}=\phi _a\left[\left(D_\mu \right)_{ab}\left(D^\mu \right)_{bc}+m^2\delta _{ab}\right]\phi _c$$
(64)
Now, utilizing LSZ reduction formulas we shall obtain the same formulas as Eqs. (53)-(LABEL:ec63) excluding appearance additional group index $`a`$ at all operators and at solution of free equations, $`f_{n_k}^{}`$, in Eqs. (51) and (52). Therefore, the proof of equivalence of the physical matrix elements of $`S`$-matrix for spin-$`0`$ particles and for GF of any number of the Yang-Milles particles can be carried out with help of the same arguments as before.
## 4 Conclusion
1. Starting from Lagrangian approach to DKP theory of spin-$`0`$ particles we constructed generating functional for many particles GF of the theory (see Eqs. (1), (3), (29)) and utilizing the LSZ reduction formulas, we strictly proved the equivalence between physical matrix elements of $`S`$-matrix in DKP and KGF theories, being the general formula (LABEL:ec63) can be applied to any type of interaction on the both theories, and consequantly to interaction spin-$`0`$ particles with quantized Maxwell, Yang Milles fields and with external gravitational fields.We also proved (see for instance Eqs. (2)-(2)) that in DKP and KGF theories the many particles GF of photons and Yang-Mills particles exactly coincide (not only physical matrix elements).
2. Considering the interaction of DKP spin-$`0`$ particles with external gravitational field we restricted by two cases: without torsion, when connection $`\mathrm{\Gamma }_{\mu \nu }^\lambda =\stackrel{}{\mathrm{\Gamma }}_{\mu \nu }^\lambda `$(see Eq. (14)) is Levi-Civita one and with torsion, when the total connection $`\mathrm{\Gamma }_{\mu \nu }^\lambda `$ is Cartan connection<sup>9</sup><sup>9</sup>9One notes that in this case the โmetric postulateโ $`_\mu g_{\nu \lambda }=0`$ does not implement., Eq. (16), antisymmetrical part of which, Eq. (18), expressed through torsion tensor. This case is one of teleparalled description of gravity . We would like to note that there is the third case interaction with gravitational field, which leads to Einstein-Cartan gravity . In the last case the metric postulate is implemented:
$$_\mu g_{\nu \lambda }=_\mu g_{\nu \lambda }\mathrm{\Gamma }_{\mu \nu }^\rho g_{\rho \lambda }\mathrm{\Gamma }_{\mu \lambda }^\rho g_{\nu \rho }=0$$
(65)
here $`\mathrm{\Gamma }_{\mu \nu }^\rho =\stackrel{}{\mathrm{\Gamma }}_{\mu \nu }^\rho +^{()}\mathrm{\Gamma }_{\mu \nu }^\rho `$
where $`{}_{}{}^{()}\mathrm{\Gamma }_{\mu \nu }^{\lambda }`$ is antisymmetrical part of Einstein-Cartan connection. The equation for $`\psi `$ now has the view (instead(19)):
$$\left(i\beta ^\mu \left(_\mu +^{()}\mathrm{\Gamma }_{\mu \rho }^\rho \right)m\right)\psi =0$$
(66)
The proof of the equivalence of DKP and KGF equations is the same.
3. For unstable particles we suggest the following fenomenological Lagrangian of interaction:
$$_{int}=\lambda \left(\overline{\psi }^M\beta _\mu \psi ^m+\overline{\psi }^m\beta _\mu \psi ^M\right)j^\mu \left(x\right)$$
(67)
where $`\psi ^M,`$ $`\psi ^m`$ are the DKP functions, decribing accordingly $`K_L^0`$ and $`\pi `$ mesons with masses $`M`$ and $`m`$; $`j_\mu =\left(\overline{e}\gamma _5\gamma _\mu \nu \right)+\left(\overline{\nu }\gamma _5\gamma _\mu e\right)`$; $`\lambda `$-fenomenological constant of point-like interaction.
Then it is easy to show in $`\lambda ^2`$-approximation that the imaginary part of GF of $`K_L^0`$ meson, which determines the amplitude of probability decay $`K_L^0\pi ^++e^{}+\overline{\nu }`$, in the DKP theory, concides with that of in KGF theories.
## 5 Acknowledgments
V. Ya. F. thanks to FAPESP for support (grant number 98/06237-0) and RFFI for partial support (grant number 99-01-00376). B. M. P. thanks to CNPq for partial support. |
warning/0003/astro-ph0003117.html | ar5iv | text | # MHD Models of Axisymmetric Protostellar Jets
## 1 Introduction
The most promising mechanism for the production of supersonic, highly collimated jets from low mass young stellar objects is by magnetic forces associated either directly with an accretion disk (Kรถnigl & Pudritz 1999), or with the interaction between an accretion disk and a magnetized central star (Shu et al. 1999). Magnetic fields are also thought to contribute to the collimation of the jets on larger scales, (although only slowly so that the observed jet may only be a part of a much broader wind, e.g., Ostriker 1997). Thus, unless the outflowing material is highly resistive (which seems unlikely), protostellar jets should contain a dynamically important magnetic field which may affect both the propagation and stability of the outflow. Observation of the magnetic field strength associated with protostellar jets is difficult. However, by fitting one-dimensional radiative shock models to the observed line ratios in the bow shock of HH47, Morse et al. (1993) inferred an upper limit to the magnetic field in the ambient medium upstream of the jet of $`30\mu `$G, a value which they argued was too small to affect the dynamics except by increasing the cooling lengths behind radiative shocks. More recently, Ray et al. (1997) have reported direct evidence for strong fields in the outflow associated with the source T Tau S through the detection of opposite circular polarization in the two spatially resolved outflow lobes. These authors infer very high field strengths within the lobes; several Gauss at a distance of few tens of AU from the central star.
If protostellar jets contain strong magnetic fields as expected from theory, there may be a signature of such fields in their dynamics. While hydrodynamical studies of cooling, dense protostellar jets are widely available in the literature (e.g., see Raga 1995; Cabrit 1997 for reviews), MHD models are less well explored. MHD studies of extragalactic (i.e., underdense and non-radiative) jets have been reported by Clarke, Norman, & Burns (1986), and Lind et al. (1989) for the case of toroidal fields, and by Kรถssl et al. (1990a; 1990b) for the case of toroidal and axial fields (see also Clarke 1996 for a recent review). For strong fields, characterized by a small value of $`\beta 8\pi P/B^2`$ where B is the field strength and P the thermal pressure (although note such fields may still have an energy density small compared to the kinetic energy density in the flow), it is found that the cocoon formed by lateral expansion of hot, shocked gas from the head of the jet is strongly inhibited. Instead, a magnetically confined โnose coneโ of shock processed jet material is formed between the bow shock and Mach disk. In addition, the stability properties of the jet beam itself are strongly altered by the presence of a magnetic field. MHD studies of the propagation of overdense, non-radiative jets into a uniform ambient medium which contains a helical magnetic field everywhere have been reported by Todo et al. (1992). They describe results for a number of field strengths and approximate the effects of radiative cooling with an equation of state with $`\gamma =1.2`$. The formation of a nose cone, as well as suppression of the cocoon, is evident for strong fields. In subsequent three-dimensional simulations (Todo et al. 1993), it was found that a magnetic field strength of $`70\mu `$G was sufficient to produce nonaxisymmetric kink instabilities in the jet beam which the authors suggest may be relevant to bends observed in the jets from some protostellar systems. This field strength is approaching the value inferred by Morse et al. (1993).
More recently, several authors have reported results from MHD studies of dense jets in which optically thin radiative cooling is directly coupled to the dynamics. Frank et al. (1998, hereafter FRJN-C) report axisymmetric studies which confirm the formation of nose cones, and (for the particular form of the initial field distribution adopted in the jet) the presence of pinch modes driven by magnetic tension. Cerqueira et al. (1997, hereafter CGH; see also de Gouveia Dal Pino & Cerqueira 1996) have used an SPH code which has been extended with an algorithm designed to represent magnetic forces to study the propagation of dense, cooling jets in three-dimensions. They find that the fragmentation of the dense shell formed by cooling at the head of the jet is strongly affected by magnetic forces, and they use these results to conclude that the clumpy structure observed in jets provides evidence that near the head of the jet the field must be primarily axial rather than helical. In their latest simulations (Cerqueira & de Gouveia Dal Pino 1999, hereafter CG), these authors also report evidence for both modes of the MHD Kelvin-Helmholtz (KH) and a current-driven pinch instability in models of protostellar jets. Finally, Frank et al. (1999a; see also Lery & Frank 1999) have studied the propagation of MHD jets launched by Keplerian accretion disks using analytic models for magneto-centrifugal outflows to describe the jet structure at the nozzle. A variety of complex behavior is seen in this case, some of which results from MHD effects and some which results from the non-uniform radial structure of the jet beam.
Given the apparent sensitivity of jet dynamics to the assumed initial orientation and strength of the magnetic field within the jet beam suggested by published numerical results, it is clearly useful to survey different initial field profiles. In this paper, we present the results from a large number of two-dimensional, axisymmetric MHD simulations of dense, cooling jets. To emphasize the effects produced solely by the magnetic field, we study jets with a uniform (โtop-hatโ) radial profile propagating into an initially uniform ambient medium. Since the theory of the production of such jets suggests a primarily toroidal field within the outflow far from the source, we focus on toroidal fields with different strengths and radial profiles. In most of our models, the ambient medium into which the jet propagates is unmagnetized, although in some cases we add a poloidal field which threads both the jet and ambient gas initially. Given the importance of nonaxisymmetric instabilities to the dynamics of magnetized jets (Todo et al. 1993), fully three-dimensional MHD simulations of cooling jets are important (CGH; CG), however in order to cover parameter space efficiently we confine the models described in this study to axisymmetry. Moreover, for the highly supermagnetosonic jets studied here, non-axisymmetric modes of, e.g., the K-H instability grow slowly and will not affect the jet in regions close to the nozzle. Of course, precession of the jet beam, a clumpy ambient medium, or a much lower magnetosonic Mach number could introduce three-dimensional effects much earlier. Our results extend previous studies in that we evolve the jet for much longer (using a grid roughly three times larger than used by FRJN-C or CG), and we focus our attention on strongly magnetized jet beams propagating into an unmagnetized ambient medium. Where there is overlap, we compare and contrast our results to previous studies throughout this paper.
The paper is organized as follows. We describe our numerical techniques in ยง2. As an aid to the interpretation of our results, we review the relevant stability properties of MHD jets in ยง3. Our results for steady are presented in ยง4, and for pulsed jets in ยง5. In ยง6 we summarize and conclude.
## 2 Method
### 2.1 MHD Algorithms
The dynamical evolution of a non-relativistic magnetized jet is given by solutions to the equations of MHD:
$$\frac{D\rho }{Dt}+\rho ๐ฏ=0,$$
(1)
$$\rho \frac{D๐ฏ}{Dt}=p+(\times ๐\times ๐)/4\pi $$
(2)
$$\rho \frac{D}{Dt}\frac{e}{\rho }=p๐ฏ\mathrm{\Lambda }+H,$$
(3)
$$\frac{๐}{t}=\times (๐ฏ\times ๐).$$
(4)
In these equations, $`D/Dt=/t+๐ฏ`$ is the convective derivative, $`\mathrm{\Lambda }`$ and H represent cooling and heating terms, respectively, and the other symbols have their usual meanings. We assume an ideal gas law for the equation of state; that is, $`p=(\gamma 1)e`$, where $`\gamma `$ = 5/3, and $`e`$ is the internal energy density. We adopt the assumptions of ideal MHD in writing equations (1) โ (4), in particular we assume the magnetic field is perfectly coupled to the fluid. While temperatures in the jet beam and in the strong shocks at the head of the jet are probably high enough (and therefore the gas is ionized enough) to keep the field well coupled on sub-parsec scales (Frank et al 1999b), it is possible that in slower shocks near the wings of the outflow, non-ideal MHD effects may become important, especially in dense ambient gas. However, the study of non-ideal MHD effects (for example, the structure of โC-typeโ shocks, Draine & McKee 1993) in the outflow is beyond the scope of this paper.
The second and third terms on the right-hand-side of equation (3) represent energy loss via optically thin cooling and heating respectively. An accurate representation of these terms is one of the most important, and yet challenging, aspects of dynamical studies of cooling jets. Previous work has shown significant differences in the evolution of cooling jets using coronal versus non-equilibrium cooling rates (Stone & Norman 1993). Recently, Raga et al. (1997) have developed an accurate and efficient formulation to approximate non-equilibrium cooling rates in interstellar gas. However, given that the focus of this paper is the effect of magnetic fields on the propagation of cooling jets, we have chosen to adopt only a simple formulation for the optically thin cooling rate: the coronal cooling curve for interstellar gas calculated by Dalgarno & McCray (1972). We allow gas to cool to a temperature of 100 K, below this we set the cooling rate to zero.
We use the ZEUS code (Stone & Norman 1992a; 1992b) to solve the MHD equations in cylindrical coordinates assuming axisymmetry. The stiff cooling and heating terms are differenced implicitly, resulting in a nonlinear equation at every grid point which is solved using a Newton-Raphson iteration scheme. This step is operator split from the rest of the MHD equations. No other additions or modifications to the ZEUS algorithm were required.
### 2.2 Initial and Boundary Conditions
The axisymmetric simulations are performed on a grid of size $`0r20R_j`$ in the transverse direction, and $`0z100R_j`$ in the axial direction, where $`R_j`$ is the initial jet radius. Except where indicated we use 400 uniform zones in the transverse direction and 2000 uniform zones in the axial direction, so that 20 zones span the initial jet radius. Reflecting boundary conditions are used along the inner $`r`$ boundary, while outflow boundary conditions are used along the outer $`r`$, outer $`z`$, and inner $`z`$ boundary for $`r>R_j`$. For $`rR_j`$ along the inner $`z`$ boundary, inflow boundary conditions are used with the variables held fixed at the values appropriate to the initial equilibrium structure of the jet.
Initially the jet is assumed to be perfectly collimated. Motivated by observations of various protostellar jets (e.g., Ray 1996), we adopt the following values as characteristic of observed systems: an initial jet radius $`R_j=2.5\times 10^{15}`$ cm, a uniform density $`n_j=1000`$ cm<sup>-3</sup>, and a uniform jet velocity $`v_j=332`$ km s<sup>-1</sup>. The gas pressure at the surface of the jet is taken to be $`p_0=1.38\times 10^{10}`$ dyne cm<sup>-2</sup>, corresponding to a temperature of $`T_j=10^3`$ K and sound speed $`a_j=5.25`$ km s<sup>-1</sup>. As described below, in order to ensure magnetostatic equilibrium in the radial direction, in some simulations we adopt a radially varying pressure profile. The ambient medium is stationary, with uniform density $`n_a=100`$ cm<sup>-3</sup> (so that the jet is overdense, i.e. $`\eta n_j/n_a=10`$), and temperature $`T_a=50`$ K. The sound speed in the ambient gas is $`a_a=1.17`$ km s<sup>-1</sup>. Note that with these choices, the jet is overpressured with respect to the ambient medium, i.e., $`p_ap_0/\alpha `$ and $`\alpha =200`$. Several purely adiabatic simulations were performed for comparison with the cooling models, in this case we have taken $`p_a=p_0`$, $`T_a=10,000`$ K and $`a_a=16.6`$ km s<sup>-1</sup>, and the jet is initially in pressure equilibrium with the ambient medium. Physically our radiatively cooling models represent a dense jet beam surrounded by a hotter cocoon embedded in a cold ISM. The kinetic model of a dense jet embedded in a less dense cocoon which is in turn ensheathed by a dense and cold molecular gas is consistent with detailed observation of a number of protostellar jet systems including HH111 (Nagar et al. 1997).
Since the strength and orientation of the magnetic field in protostellar jets is not determined observationally, there is considerable freedom in specifying the initial field configuration. We choose to focus on primarily toroidal fields in this paper (although we do present models which include both toroidal and axial fields) as theoretical models of magnetically driven outflows predict the field should become primarily toroidal asymptotically. In the absence of a definitive theory for the radial profile of the magnetic field, we assume the toroidal magnetic field is zero on the jet axis, achieves a maximum at some radial position in the jet interior $`r_m`$, and returns to zero at the outer jet boundary. This implies that all currents which support the field are confined within the jet (although as the simulation progresses magnetic field, and therefore current, is convected into the cocoon). Within the jet, we adopt the simple profile for the toroidal magnetic field
$$B_\varphi (r)=\{\begin{array}{cc}B_{\varphi ,m}\frac{r}{r_m}& 0rr_m\\ B_{\varphi ,m}\frac{R_jr}{R_jr_m}& r_mrR_j\\ 0& R_j<r\end{array}$$
(5)
This profile is identical to that used by Lind et al. (1989) and by FRJN-C when $`0rr_m`$ and when $`r>R_j`$ but differs for $`r_mrR_j`$ where they used a force-free $`B_\varphi (r)=B_{\varphi ,m}(r_m/r)`$ (which requires a large return current located at the jet surface). CG have studied both longitudinal and helical fields of the form used by Todo et al. (1992) throughout the computational domain. Our toroidal magnetic field corresponds to a uniform current density inside radius $`r_m`$, and a distributed return current in the outer portion of the jet. In adiabatic and cooling jet simulations designed to be compared directly, the toroidal magnetic field reaches a maximum value at $`r_m/R_j=0.9`$ with $`\beta _1\beta _m=8\pi p_0/B_{\varphi ,m}^2=1`$ ($`B_{\varphi ,m}=58.8`$ $`\mu G`$) and with $`\beta _{1/4}\beta _m=0.25`$ ($`B_{\varphi ,m}=117.6`$ $`\mu G`$). To illustrate the toroidal magnetic field profile adopted here, we plot equation (5) for $`\beta _m=1`$ and $`r_m=0.9`$ in Figure 1.
The equation of hydromagnetic equilibrium
$$\frac{d}{dr}\left(p_j(r)+\frac{B_z^2(r)}{8\pi }+\frac{B_\varphi ^2(r)}{8\pi }\right)=\frac{B_\varphi ^2(r)}{4\pi r},$$
(6)
has been used to establish a suitable radial gas pressure profile in the jet
$$p_j(r)=\{\begin{array}{cc}\left\{2\left[1(r/r_m)^2\right]\frac{B_{\varphi ,m}^2}{8\pi p_0}+p_m/p_0\right\}p_0& 0rr_m\\ \left\{1\frac{2}{(1r_m/R_j)^2}\left[3(1r/R_j)(1r^2/R_j^2)+\mathrm{ln}(r/R_j)\right]\frac{B_{\varphi ,m}^2}{8\pi p_0}\right\}p_0& r_mrR_j\\ p_0& r=R_j\end{array}$$
(7)
where $`p_m=p_j(r_m)`$ and
$$p_j(r_m)=p_0\frac{2}{(1r_m/R_j)^2}\left[3(1r_m/R_j)(1r_m^2/R_j^2)+\mathrm{ln}(r_m/R_j)\right]\frac{B_{\varphi ,m}^2}{8\pi }.$$
(8)
The radial pressure profile which gives a magnetostatic equilibrium is shown in Figure 1 where we plot equation (7) for $`\beta _m=1`$ and $`r_m=0.9`$. Since the jet density is held constant, $`T_j`$ has the same functional form as $`p_j`$. In some of our simulations with strong toroidal fields, the condition of magnetohydrostatic equilibrium within the jet beam requires negative pressures in the neighborhood of $`r_m`$. In these cases we simply fix $`p_m=0`$, and scale the pressure profile given by equation (7) by an appropriate amount so that this condition is met. Of course, this means that the jet beam is no longer in strict radial equilibrium.
The profile adopted for the magnetic field is an important difference between this work and previous studies (FRJN-C, CG). To allow us to study the effect of the field geometry, we have also performed a few simulations in which the field contains a uniform axial component as well as the toroidal distribution given above, making the field helical. We will also describe the evolution of force-free fields with $`B_\varphi 1/r`$. We find that the detailed structure of a magnetized radiative jet is sensitive to the precise field geometry, thus uncertainty about this geometry is a serious limitation to modeling observed systems.
To construct models of pulsed jets, we apply a purely sinusoidal perturbation in time so the axial component of the velocity is
$$u_z(x=0)=v_j(1+A\mathrm{sin}\omega t),$$
(9)
where we choose $`A=0.25`$, and for all of the simulations $`\omega =25a_0/R_j`$ where $`a_0=16.6`$ km s<sup>-1</sup> is a fiducial reference speed (equivalent to the sound speed at a temperature of $`10^4`$ K).
The important dimensionless parameters which characterize the evolution of the jet are the density ratio $`\eta `$ (taken to be 10 for all models presented here), and the internal magnetosonic Mach number $`M_{j,ms}=v_j/a_{ms}`$, where $`a_{ms}(a_j^2+V_A^2)^{1/2}`$ and $`V_A`$ is the Alfvรฉn speed. In simulations without toroidal magnetic field or with force-free toroidal magnetic field the pressure, density, and temperature are uniform across the jet, so that the internal sonic Mach number $`M_j=v_j/a_j`$ is constant. However, a radially varying magnetic pressure requires a radially varying thermal pressure in order to initialize a jet beam in radial force balance and the sonic and magnetosonic Mach numbers are not constant. In general, we characterize all our models by the magnetosonic Mach number on the jet axis. However, since internal dynamics and timescales involve wave propagation across the jet through a medium of varying magnetosonic speed we also define linear radial averages as, for example,
$$\overline{M}_{j,ms}\frac{1}{R_j}_0^{R_j}M_{j,ms}(r)๐r.$$
(10)
### 2.3 Tests of the Method
Tests of the MHD algorithms used here have been reported in a number of papers (Stone & Norman 1992a, 1992b; Stone et al. 1992; Hawley & Stone 1995). We have also confirmed that we are able to reproduce the results of previous MHD studies of jets. In particular, we have performed simulations of adiabatic, underdense, magnetized jets using parameters identical to Clarke, Norman, & Burns (1986); we find excellent agreement between our results with those reported by these authors. In particular, we find production of a nose cone with a similar aspect ratio, strong pinch modes within the jet beam due to lack of initial radial equilibrium, and strong suppression of vortical motion in the cocoon by magnetic stresses.
## 3 Jet stability
Before describing the results of our simulations, let us review the basic stability properties of supermagnetosonic jets in order to aid interpretation. The presence of toroidal magnetic fields and the poloidal current can give rise to current driven pinch modes in addition to velocity shear driven Kelvin-Helmholtz pinch modes. In the absence of any zero order velocity shear, e.g., well inside the velocity shear surface at the jet boundary, Begelman (1998) has shown that a beam containing a toroidal magnetic field can be unstable to current driven pinch modes when
$$\frac{d\mathrm{ln}B}{d\mathrm{ln}r}>\frac{\mathrm{\Gamma }\beta (r)2}{\mathrm{\Gamma }\beta (r)+2}$$
(11)
where the plasma beta $`\beta (r)8\pi p(r)/B^2(r)`$, and $`\mathrm{\Gamma }`$ is the adiabatic index. In general, shorter pinch lengths will grow faster. If the beam contains a toroidal magnetic field and a weaker uniform axial magnetic field the condition becomes
$$\frac{d\mathrm{ln}B}{d\mathrm{ln}r}>1+\frac{1}{2}(kr\frac{B_z}{B_\varphi (r)})^2,$$
(12)
where $`k`$ is the wavenumber of the mode; in general, shorter pinch lengths, i.e., larger $`k`$, are stabilized.
The velocity driven Kelvin-Helmholtz unstable pinch body modes might appear near resonances which, in the absence of radial variation in parameters in the jet and in the external medium, are given by (e.g., Hardee, Clarke, & Rosen 1997; Hardee & Stone 1997)
$$\omega R_j/a_c(2m+1/2)\pi /2$$
()
$$\lambda /R_j\frac{2\pi }{\omega R_j/a_c}\frac{M_{j,ms}}{1+M_{j,ms}/M_c}$$
()
where $`\omega `$, $`m`$, and $`\lambda `$ are the angular frequency, body mode number $`m=`$ 1, 2, 3, etc., and wavelength, respectively. In the expressions above $`M_cv_j/a_c`$ and $`a_c`$ is the sound speed in the cocoon medium immediately outside the jet and not the sound speed in the undisturbed ambient medium. At these resonances the pinch modes grow with spatial growth rate
$$k_I(2M_{j,ms}R_j)^1\mathrm{ln}\left(4\omega R_j/a_c\right),$$
()
where $`k_I`$ is the imaginary part of the wavenumber. The spatial growth length over which a linear perturbation increases in amplitude by a factor $`e`$ is $`\mathrm{}=\left|k_I^1\right|M_{j,ms}R_j`$. On the high Mach number axisymmetric flows that we simulate here these Kelvin-Helmholtz modes grow relatively slowly.
## 4 Steady Jet Simulations
Simulation parameters for the steady jet models discussed here are listed in Table 1. In what follows, we report times in units of a sound crossing time of the jet beam, $`\tau _0R_j/a_j152`$ yr where $`a_j=5.25`$ km s<sup>-1</sup>. Magnetosonic crossing times of the jet beam are in the range, $`\tau _{ms}R_j/\overline{a}_{ms}15058`$ yr where $`\overline{a}_{ms}`$ is the linear average of the magnetosonic speed. A grid crossing time can be defined by $`\tau _gL_{grid}/v_h314`$ yr, where $`v_h=\left\{\eta ^{1/2}/\left[1+\eta ^{1/2}\right]\right\}v_j250`$ km s<sup>-1</sup> is the velocity of advance of the head of a cold (thermal and magnetic pressure much less than the ram pressure) jet. We stop the evolution when the bow shock has reached $`100R_j`$. The spatial evolution of the jetโs transverse structure is then evaluated at an axial distance of 68.5R<sub>j</sub> (a location behind the region strongly affected by jet head vortices), and the spatial evolution of the cocoonโs structure can be evaluated through comparison of cocoon properties at this distance and near to the inlet.
### 4.1 Magnetized Adiabatic Jets
In order to understand the dynamics of our magnetized radiative jets, it is necessary to first compute non-radiative jet models with an identical magnetic field profile. We have computed three models: a purely hydrodynamical jet ($`\beta _{\mathrm{}}^{Ad}`$), a weak field jet ($`\beta _1^{Ad}`$), and a strong field jet ($`\beta _{1/4}^{Ad}`$). Table 1 gives additional parameter values adopted for each simulation. Unlike our radiatively cooling jet models, all three adiabatic models are initially in pressure equilibrium with the ambient medium; i.e. we adopt $`T_a=10^4`$ K rather than the $`T_a=50`$ K used for the cooling jets. Figure 2 plots the logarithm of the density at the end of each of these three simulations. Figure 3 plots the toroidal magnetic field strength $`B_\varphi `$ at the same times for the two magnetized jet models. Note only the final $`50R_j`$ of the computational domain is shown in both figures for clarity.
Comparison of the final structure of each jet shows remarkably little difference between these models. In the purely hydrodynamical jet simulation $`\beta _{\mathrm{}}^{Ad}`$ the jet remains well collimated within its initial radius inside a low density cocoon of typical density one-half the ambient and one-twentieth the jet density, i.e., n$`{}_{c}{}^{}50`$ cm<sup>-3</sup>, and has a typical radius $`10R_j`$.
The weak toroidally magnetized jet simulation, $`\beta _1^{Ad}`$, experiences some slight collimation from the toroidal magnetic field so that by an axial distance of 68.5 R<sub>j</sub>, where r$`{}_{j}{}^{}R_j`$, the toroidal field maximum has increased by about 7% as it has moved inwards with the contracted jet. However, unlike the $`\beta _{\mathrm{}}^{Ad}`$ jet we find three significant pressure pulses with amplitudes up to twice p$`{}_{j}{}^{}(r)`$ and spacing of about 15R<sub>j</sub> (with the first pulse at an axial distance of $`30R_j`$) in the jet interior when $`0.1<r/r_j<0.6`$. The amplitudes are largest at $`r0.3r_j`$. This location is not consistent with pressure perturbations produced by the Kelvin-Helmholtz instability at the jet surface. The confinement of these pressure perturbations to the jet interior is consistent with a current driven instability, moreover this region satisfies the condition for instability (eq. ) found by Begelman (1998).
The strong toroidally magnetized jet simulation, $`\beta _{1/4}^{Ad}`$, cannot be introduced at the inlet with a true equilibrium pressure distribution: this would require a negative minimum jet pressure. As a result the jet spine expands to $`r_j1.5R_j`$ at an axial distance of $`68.5R_j`$ and then recollimates somewhat under the influence of higher external pressure near to the jet head and the stronger toroidal magnetic field. At axial distances between $`8095R_j`$ there is evidence for pressure pulses associated with pinching in the jet interior at $`r<0.2r_j`$. Three main pressure pulses with spacing of $`6R_j`$ and substructure at one-third this spacing might be due to current driven instability. Neither of the toroidally magnetized adiabatic jet simulations show evidence for a magnetically confined โnose coneโ (Clarke, Norman, & Burns 1986; Lind et al. 1989).
### 4.2 Magnetized Cooling Jets
Images from three cooling jet simulations, a purely hydrodynamical model ($`\beta _{\mathrm{}}^{Cl}`$), a weak field model ($`\beta _1^{Cl}`$), and a strong field model ($`\beta _{1/4}^{Cl}`$) with magnetic field configuration identical to that used in the three adiabatic jet simulations described in ยง4.1 are shown in Figure 4. Figure 5 shows the toroidal magnetic field strength $`B_\varphi `$ at the same times for the two magnetized jet models. Note only the final $`50R_j`$ of the computational domain is shown in both cases. The purely hydrodynamical jet simulation $`\beta _{\mathrm{}}^{Cl}`$ is directly comparable to the results presented in Stone & Norman (1993; 1994); the primary difference being the use here of the coronal cooling curve of Dalgarno & McCray (1972), as opposed to the non-equilibrium cooling rates used previously.
Unlike the adiabatic jets discussed above, the addition of a toroidal magnetic field to radiatively cooling jets leads to noticeable differences relative to the non-magnetized $`\beta _{\mathrm{}}^{Cl}`$ simulation. For example, in $`\beta _1^{Cl}`$ the jet beam has expanded in width, and the cocoon is much more uniform and smaller relative to $`\beta _{\mathrm{}}^{Cl}`$. Jet expansion in $`\beta _1^{Cl}`$ has resulted in a 50% decrease in the toroidal magnetic field strength in the jet beam. No nose cone is evident in either the $`\beta _{\mathrm{}}^{Cl}`$ or $`\beta _1^{Cl}`$ simulations.
The $`\beta _{1/4}^{Cl}`$ simulation is strongly influenced by expansion of the jet resulting from the non-equilibrium initial configuration at the inlet. Figure 6 plots radial profiles of the density, axial velocity, temperature, and toroidal magnetic field strength at an axial distance of 68.5 R<sub>j</sub> from the jet nozzle in all three simulations. From the figure, the structure of the $`\beta _{1/4}^{Cl}`$ jet can be seen to consist of an expanded spine with r$`{}_{j}{}^{}`$ 3.5R<sub>j</sub> surrounded by a sheath of thickness $``$ 3.5R<sub>j</sub>. The magnetic profile shows an approximate linear increase from the axis to 2R<sub>j</sub> and plateau to 4R<sub>j</sub> with B$`{}_{\varphi }{}^{}21`$ $`\mu G`$. At the same time, the jet temperature drops from 2,000 K on axis to a low of $``$ 100 K in the outer portion of the jet, and then jumps up to $``$ 2,000 K in the cocoon.
The stronger magnetic field along with jet expansion have resulted in significant morphological changes in jet and cocoon structure beyond an axial distance of about 70 R<sub>j</sub> in $`\beta _{1/4}^{Cl}`$. At this distance the jet velocity suddenly decreases by more than a factor of two at an โupstream shockโ and a nose cone of length $`25R_j`$ is โpushedโ ahead of a โdownstream shockโ. The nose cone drives a more conventional jet terminal and bow shock. The temporal evolution leading to this structure is shown in Figure 7, which plots the logarithm of the density at 5 times during the simulation, $`(0.4,0.8,1.2,1.6,1.8)\times \tau _0(61,122,182,243,274)`$ yr. Initially the jet remains well collimated. Strong cooling at the head of the jet results in most of the shocked ambient and jet gas collecting in a dense shell ahead of the Mach disk. Once the jet has propagated about $`20R_j`$, it begins to expand due to lack of magnetohydrostatic balance. Consequently, the head of the jet grows laterally, and becomes filamentary. Most of the shock processed gas is confined to a loosely defined โnose coneโ ahead of the Mach disk. Detailed structure at the head of the jet is shown further in Figure 8, which plots the logarithm of the temperature, the divergence of the velocity (an indicator of shocks), and velocity vectors at the end of the simulation. Comparison of the temperature and velocity divergence plots clearly shows that the highest temperatures are located in filaments immediately behind strong shocks in the flow, such as the Mach disk and outer bow shock. Very little non-axial motion is observed in the velocity vectors.
### 4.3 Cooling Jets with Different Magnetic Field Profiles
Three additional magnetic field profiles have been studied for steady radiatively cooling jets. One simulation, $`\beta _{ff}^{Cl}`$, has a toroidal force-free field in the jet with $`B_\varphi 1/r`$. In this simulation the maximum field strength $`B_{\varphi ,m}=236.3`$ $`\mu G`$ ($`\beta _m=0.062`$) at $`r_m/R_j=0.05`$ (the innermost computational zone) and the magnetic field is set to zero at the jet surface. Within the force-free jet, temperature and thermal pressure are constant. In a second simulation, $`\beta _{Rm}^{Cl}`$, the magnetic and pressure profiles given by equations (5) and (7) are used with $`\beta _m=0.25`$ ($`B_{\varphi ,m}=117.7`$ $`\mu G`$) at $`r_m/R_j=0.2`$. Unlike the $`\beta _{1/4}^{Cl}`$ simulation, in this simulation a true equilibrium pressure profile can be achieved and outside the toroidal field maximum the field configuration is not too different from the force free simulation $`\beta _{ff}^{Cl}`$. Finally, a third simulation, $`\beta _{\varphi z}^{Cl}`$, includes an axial magnetic field along with a toroidal magnetic field. In this simulation a constant axial magnetic field in the jet and in the ambient medium of strength $`B_z=58.8`$ $`\mu G`$ ($`\beta _z=1`$) has been added to a toroidal magnetic field profile given by equation (5) with, $`r_m/R_j=0.8`$ and $`\beta _m=0.25`$ ($`B_{\varphi ,m}=117.7`$ $`\mu G`$). In this simulation, the pressure profile in the jet is similar to that in the $`\beta _{1/4}^{Cl}`$ simulation and is not in equilibrium. Figure 9 plots the logarithm of the density at the final time for each of these simulations (along with the $`\beta _{1/4}^{Cl}`$ simulation shown in Figures 7 & 8), and Figure 10 plots the corresponding toroidal magnetic field strength. Radial profiles of a variety of quantities in the jet beam for each simulation at an axial distance of 68.5R<sub>j</sub> are given in Figure 6.
In $`\beta _{ff}^{Cl}`$ the jet cools radiatively and the axial velocity remains constant out to $``$ 90 R<sub>j</sub> where the velocity abruptly drops by about 20% at a terminal โupstreamโ shock. A modest nose cone of material about 9 R<sub>j</sub> in length precedes a โdownstreamโ shock and drives a terminal and bow shock. The jet expands as the initial jet pressure is above the radiatively cooled cocoon pressure at the inlet. The radial structure (Figure 6) is similar in dimension to that found in the $`\beta _{\mathrm{}}^{Cl}`$ and $`\beta _1^{Cl}`$ simulations, although the cocoon size is slightly less than in the $`\beta _1^{Cl}`$ simulation.
The $`\beta _{Rm}^{Cl}`$ simulation serves as an intermediate case between the force-free $`\beta _{ff}^{Cl}`$ simulation and the non force-free $`\beta _{1/4}^{Cl}`$ simulation, and can also be compared to the $`\beta _1^{Cl}`$ simulation. Inside the jet when $`r>0.2r_j`$ the magnetic field profile is very similar to the force-free case. At the inlet the jet is overpressured relative to the radiatively cooled cocoon but jet expansion does not begin until an axial distance of about 40 R<sub>j</sub>. Interior to this distance nine well defined pressure oscillations with a spacing of $``$ 3.3 R<sub>j</sub> are observed at jet radii $`r<0.15r_j`$, and probably are the result of a current driven pinch mode. Pressure fluctuations attributable to this pinch mode are on the order of $`\pm 2\%3\%`$ and these fluctuations disappear as the jet begins to expand. At an axial distance of 68.5 R<sub>j</sub> (Figure 6) the jet spine and sheath are reduced in size by about 25% relative to the force-free configuration. The jet is in approximate pressure balance with the cocoon at this point. We note that the jet spine contains a relatively high pressure and temperature central region confined by the magnetic field. The cocoon is also reduced in size by about 25% relative to the force-free configuration. A modest nose cone about 10 R<sub>j</sub> long leads the jet in this case after a velocity drop of 15% in an โupstreamโ shock.
Finally in the $`\beta _{\varphi z}^{Cl}`$ simulation we consider the effect of adding an axial magnetic field to the toroidal magnetic field configuration used in $`\beta _{1/4}^{Cl}`$. The axial field imposed over the entire computational grid has served to ameliorate the large jet expansion observed in $`\beta _{1/4}^{Cl}`$ and at an axial distance of 68.5 R<sub>j</sub> (see Figure 6) the jet spine r$`{}_{j}{}^{}3R_j`$ and sheath of thickness $`R_j`$ together equal the jet spine width in $`\beta _{1/4}^{Cl}`$. This width is similar to that seen for the $`\beta _{ff}^{Cl}`$ jet and in this simulation the jet is at lower pressure than the hotter cocoon. Toroidal magnetic field resides within the jet spine and sheath, while the axial magnetic field has been reduced within the jet by jet expansion but increased by compression in the outer part of the cocoon. This additional magnetic field and accompanying high sheath and higher cocoon density has helped to confine the jet. The presence of axial magnetic field has almost but not quite eliminated the extensive nose cone observed in the $`\beta _{1/4}^{Cl}`$ simulation.
## 5 Pulsed Jets
Some of the most prominent structural features in highly collimated jets associated with protostellar objects are the bright emission knots in the flow close to the source (see Reipurth 1997 for a recent review of the observations). It has been suggested that symmetric pinch modes of the Kelvin-Helmholtz (K-H) instability might give rise to such knots (Bรผhrke, Mundt, & Ray 1988), which has motivated detailed hydrodynamical modeling of the nonlinear evolution of symmetric K-H modes in cooling jets (Massaglia et al. 1992; Bodo et al. 1994; Rossi et al. 1997; Downes & Ray 1998). However the large proper motion observed in some cases (Eislรถffel & Mundt 1994), and direct kinematic evidence provided by spectroscopic observations (Reipurth 1997) have shown that some knots are associated with velocity variability in the outflow. Thus, while the nonlinear stage of nonaxisymmetric modes of the K-H instability might still prove relevant to understanding the helical structures or wiggles observed in some protostellar jets (Hardee & Stone 1997; Stone, Xu, & Hardee 1997), most of the internal knot structure in the jets is probably produced by flow variability in the jet beam.
The hydrodynamics of velocity-variable, or pulsed, jets has been widely studied in the literature (see Raga 1993 for a review). In an important early contribution, Raga & Kofman (1992) demonstrated that a sinusoidal velocity variation at the jet inlet steepens into a sawtooth pattern downstream. One-dimensional (Hartigan & Raymond 1993), followed by multidimensional hydrodynamical simulations (Stone & Norman 1993; de Gouveia Dal Pino & Benz 1994; Biro & Raga 1994; Biro 1996; Suttner et al. 1997) have shown that the velocity discontinuities in the sawtooth pattern consist of a shock pair: an upstream shock decelerating high velocity gas as it collides with the pulse, and a downstream shock sweeping up low velocity material ahead of the pulse. These shock pairs move apart at a constant rate causing the pulse width to grow linearly in time up to some asymptotic value. In the absence of magnetic fields this evolution can be understood by a simple analytic model (Falle & Raga 1993). Recently, there has been considerable effort in understanding how such โinternal working surfacesโ (Raga et al. 1990) affect the dynamics and observed properties of jets. In this section, we present the results of an investigation of the magnetohydrodynamics (MHD) of radiatively cooling pulsed jets.
Parameters for the simulations discussed below are listed in Table 2. Because shock compression on the downstream and upstream sides of pulses generated by velocity fluctuation is a function of jet radius, we include in Table 2 area weighted averages as, for example,
$$M_{j,ms}\frac{2}{R_j^2}_0^{R_j}M_{j,ms}(r)r๐r\text{ .}$$
(13)
### 5.1 Variation in Magnetic Field Strength
We have computed a purely hydrodynamic model with $`\beta _\varphi =\mathrm{}`$, hereafter denoted as $`\beta _{\mathrm{}}^P`$, to serve as a benchmark. Two simulations were performed using a toroidal magnetic field profile in the jet beam given by equation 5 in Paper II, with r$`{}_{m}{}^{}=0.9R_j`$. The first model has a peak field strength of $`B_{\varphi ,m}=58.8\mu G`$ corresponding to an equipartition field, i.e. $`\beta _{\varphi ,m}8\pi p_0/B_{\varphi ,m}^2=1`$, hereafter denoted as $`\beta _1^P`$. The second model has a peak field strength $`B_{\varphi ,m}=117.6\mu G`$ corresponding to a strong field where $`\beta _{\varphi ,m}=0.25`$, hereafter denoted as $`\beta _{1/4}^P`$. These profiles are identical to those adopted in the steady jet simulations $`\beta _{\mathrm{}}^{Cl}`$, $`\beta _1^{Cl}`$, and $`\beta _{1/4}^{Cl}`$.
In Figure 11 we show images of the logarithm of the density at the final time in the three simulations. The $`\beta _{\mathrm{}}^P`$ model is directly comparable to previous results of hydrodynamical simulations, e.g., Stone & Norman (1993). Despite the use of a different numerical method (a PPM algorithm was used in this earlier work, instead of the ZEUS code used here) and lower numerical resolution, there is excellent agreement in the overall structure of the jet. The steepening of the sinusoidal pulses into thin, dense sheets by $`z10R_j`$ is evident. Thereafter, the pulses are bounded by two shocks. Significant mass flux into the cocoon from the high pressure shocked gas within the pulse is evident as smooth flows. Each pulse is seen to drive a shock into the cocoon, so that the cocoon is much broader than steady jet models using the same parameters. The structure at the head of the jet is complex as pulses begin to merge with one another, and interact with the ambient medium. Note that the pulse width increases linearly with distance from the nozzle. For the purely hydrodynamic jet, the rate of increase in size is set by the balance between radial mass loss into the cocoon from the pulse โsurfaceโ and the mass flux into the pulse through isothermal shocks (Falle & Raga 1993).
As the magnetic field strength is increased (models $`\beta _1^P`$ and $`\beta _{1/4}^P`$ shown in the lower two panels of Figure 11), the gross properties of the pulsed jet remain the same. Two features of the pulses are clearly modified by the magnetic field. Firstly, the postshock density within the pulses on the axis increases with the field strength. Secondly, the rate of increase in the pulse width increases with increasing magnetic field strength. While the presence of hoop stresses associated with the toroidal field clearly influences the amount of material ejected from the pulses into the cocoon, the maximum radial distance that material is ejected into the cocoon does not decrease with field strength. In all three simulations, the wings generated by the pulses extend to a radial distance of $`5R_j`$. Figure 12 plots the logarithm of the toroidal magnetic field strength in $`\beta _1^P`$ and $`\beta _{1/4}^P`$. The field acts as a tracer of jet material since the ambient gas is unmagnetized. Compression of the magnetic field by the internal shocks associated with the pulses, and ejection of material and the magnetic field into the cocoon is evident. Note this ejection occurs over a larger area in $`\beta _{1/4}^P`$ as might be expected given the larger pulse width.
In order to illustrate the variation of properties of the pulses with magnetic field strength, we plot in Figure 13 the density, pressure, axial velocity, and toroidal magnetic field strength along an axial slice through the center of all three jets at time $`t=1.2\tau _0`$ โ note that this time is not the same as that used in Figures 11 and 12. Only a limited segment of the axial domain between $`20R_j`$ and $`45R_j`$ is shown for clarity. The plot shows three pulses developing within this region, one centered near $`23R_j`$, the next at $`31R_j`$, and the last at $`38R_j`$. In both cases the magnetic field strength declines outwards from the origin over the region shown in Figure 13 by a factor 2 โ 3. Several trends are clearly evident in the plot. Note the increase in width of the pulses with axial distance. The strongly magnetized model shows the widest pulses. Interestingly, the gas pressure is nearly identical within all three jets. The axial density is much higher in the strongly magnetized pulsed jet, increasing from 5 times (in the first pulse) to 20 times (in the last pulse) larger than the hydrodynamic pulsed jet. The axial velocity shows the classic sawtooth pattern expected for a sinusoidally pulsed jet (Raga & Kofman 1992), modified by the shock pair which appears at each step (e.g., Stone & Norman 1993). The toroidal field strength follows the density profile closely as would be expected in ideal MHD when the magnetic field is parallel to the shock interface.
In all three jets, the pulse width increases linearly with axial distance. At $`z50R_j`$, the pulse widths are about $`4R_j`$, $`5R_j`$, and $`7R_j`$ in $`\beta _{\mathrm{}}^P`$, $`\beta _1^P`$, and $`\beta _{1/4}^P`$, respectively. To zeroth order the linear growth rate in the pulse width scales with $`1/M_{j,ms}`$ rather than the sound speed (Falle & Raga 1993) (see Table 2). In $`\beta _1^P`$ and $`\beta _{\mathrm{}}^P`$ the pulses remain separated along the entire length of the jet, so that the upstream and downstream shock pair is evident nearly to the jet head. On the other hand, the spreading of the pulse width is so large in the $`\beta _{1/4}^P`$ model that the pulses merge before reaching the head of the jet, leading to the disappearance of the upstream shock โ leaving only a sequence of downstream shocks โ at $`z>60R_j`$.
The observed differences in pulse density, radial density profile and pulse width are the result of hoop stresses which confine some of the shocked jet material within the pulses to near the axis, and the result of the lack of radial pressure equilibrium in the material after it passes through the upstream and downstream shocks into the pulse. Recall that the unshocked material in $`\beta _1^P`$ is in radial pressure equilibrium. In an isothermal shock the toroidal magnetic pressure increases $`\rho ^2`$ whereas thermal pressure increases $`\rho `$. Thus, in general, the toroidal plasma $`\beta `$ behind the shocks should decrease relative to the toroidal plasma $`\beta `$ in the unshocked jet material by up to $`\rho ^1`$. Of course, the toroidal plasma $`\beta `$ in the unshocked and shocked pulse material is a function of radius and the relative importance of thermal versus magnetic pressure is further modified by material and magnetic ejection from the pulse. However, in $`\beta _{1/4}^P`$ the area weighted plasma $`\beta `$ in a typical pulse is $`\beta _\varphi 2`$ (compared to $`\beta _\varphi 8`$ in the unshocked jet material), and in this case magnetic pressure can play a significant role in supporting the pulses against ram pressure.
Considerable toroidal field accumulates at the head of the jet into a nose-cone-like structure; this is most evident in the very strong field case. For our parameters typical magnetic field strengths in the pulses at $`z>20R_j`$ are about $`50\mu `$G and $`75\mu `$G in $`\beta _1^P`$ and $`\beta _{1/4}^P`$, respectively, and the field strength declines as the distance from the inlet increases. Between the pulses the magnetic field strength is reduced from these values by over a factor of 10. Near the head of the jet in the complex interactions between shocked jet material ejected from the pulses and shocked ambient gas, field strengths are as high as $`190\mu G`$ in $`\beta _1^P`$, and $`570\mu G`$ in $`\beta _{1/4}^P`$.
### 5.2 Radial Variation in Magnetic Field and Pulse Structure
We have studied the effect of varying the geometry of the magnetic field in the jet beam using three different simulations. The first, $`\beta _1^P`$, is the model discussed above in which $`B_{\varphi ,m}=58.8\mu G`$ at $`r=0.9R_j`$. The second, $`\beta _{Rm}^P`$, is another equipartition toroidal magnetic field model but with $`B_{\varphi ,m}=58.8\mu G`$ at $`r=0.2R_j`$. The third, $`\beta _{\varphi z}^P`$, is a model with $`B_{\varphi ,m}=58.8\mu G`$ at $`r=0.8R_j`$, and which contains a uniform axial field of $`B_z=58.8\mu G`$, with $`\beta _z8\pi p_0/B_z^2=1`$ added throughout the computational domain. In all three of these simulations, the jet beam is initially in exact radial equilibrium.
Images of the logarithm of the density at the final time in each of these simulations are shown in Figure 14. A comparison between $`\beta _1^P`$ and $`\beta _{Rm}^P`$ shows that when the magnetic field is peaked close to the axis, the pulses widen non-uniformly, growing in width more rapidly near the axis of the jet than at the surface. The size and shape of the wings of the pulses is similar in both models. A strong axial magnetic field (model $`\beta _{\varphi z}^P`$) helps to confine material in the pulses and reduces the amount of material ejected to the cocoon.
Images of the logarithm of the toroidal magnetic field strength in $`\beta _1^P`$, $`\beta _{Rm}^P`$, and $`\beta _{\varphi z}^P`$ are shown in Figure 15. In $`\beta _1^P`$, considerable toroidal magnetic field has been injected into the cocoon. Since the field tracks jet material, this indicates a large amount of shocked jet gas is ejected from the pulses. On the other hand, in $`\beta _{Rm}^P`$ the strongest toroidal field is found close to the axis, and there is much less magnetic field in the cocoon, indicating reduced ejection of shocked jet material from the pulses. In $`\beta _{\varphi z}^P`$ the toroidal magnetic field is more confined towards the axis indicating that considerably less material is ejected from the pulses into the cocoon, and the cocoon is smaller in this simulation than in the other two simulations.
In Figure 16 we plot various quantities as a function of radius for pulses at an axial distance $`z/R_j=`$ 48, 53, 51, 47.5 and 51 at the final time in simulations $`\beta _{\mathrm{}}^P`$, $`\beta _1^P`$, $`\beta _{1/4}^P`$, $`\beta _{Rm}^P`$ and $`\beta _{\varphi z}^P`$, respectively. These locations are the centers of last pulse which is non-interacting with its neighbors at the final time in each simulation. Note that $`\beta _{1/4}^P`$ and $`\beta _{Rm}^P`$ result in similar axially peaked density pulse profiles (although the axial density in $`\beta _{1/4}^P`$ is four times that in $`\beta _{Rm}^P`$). Other magnetic field configurations are less axially peaked but are still very different from the โtop hatโ profile of the hydrodynamical simulations (solid line in each panel). In all cases the magnetized pulses are significantly colder on the axis, typically by more than a factor of 2, than the hydrodynamic pulse. The hydrodynamic pulse has a โtop hatโ temperature profile with $`T_{pulse}1,300`$ K. Expansion velocities evaluated at $`r=0.8R_j`$ lie within $`\pm 15\%`$ of $`a_j`$ ($`a_j=a_0/3.16`$) and are relatively independent of magnetic field strength or configuration. Thus, we might expect an equilibrium pulse width (mass into the pulse through upstream and downstream shocks balanced by transverse mass ejection) to scale approximately proportional to the density at the $`r=0.8R_j`$ surface, i.e., with the mass ejection rate per unit surface area evaluated at $`r=0.8R_j`$.
At least approximately we would expect the pulse expansion rate and ultimate width in $`\beta _1^P`$ and $`\beta _{\varphi z}^P`$ to be comparable but twice that (and in $`\beta _{1/4}^P`$ and $`\beta _{Rm}^P`$ comparable but about four times that) observed in the hydrodynamic case. Our simulations do in fact show an increase in the rate at which pulses widen consistent with this analysis, although the rate of this increase is not as fast as this simple argument would predict.
From Figure 16, the maximum radial ejection velocity from the pulses occurs at a radial distance much larger than the pulse surface at $`r=0.8R_j`$. It is affected by the magnetic field configuration and strength, with the highest velocities observed for the strongest toroidal magnetic field case, $`\beta _{1/4}^P`$, and lowest velocities observed when an axial magnetic field is present, $`\beta _{\varphi z}^P`$. Changes to the unshocked toroidal plasma $`\beta `$ in the pulses is also dependent on the magnetic field configuration and strength, and the radial variation in $`\beta _\varphi `$ shown in Figure 16 is much different from that at the inlet (see Table 2). If we characterize the effects of the magnetic field by the area weighted toroidal plasma $`\beta `$ then we find that $`\beta _\varphi _{pulse}\beta _\varphi _{jet}`$ in cases $`\beta _1^P`$ and $`\beta _{\varphi z}^P`$, whereas $`\beta _\varphi _{pulse}<<\beta _\varphi _{jet}`$ in cases $`\beta _{1/4}^P`$ and $`\beta _{Rm}^P`$. Note that this properly groups the magnetic simulations with their observed spreading rates. However, only in case $`\beta _{1/4}^P`$ where $`\beta _\varphi _{pulse}1.5`$ can there be significant magnetic pressure effects. We conclude that the differences in pulse width spreading rate are the result of the radial density profile in the pulse induced by radial pressure equilibrium with $`B_\varphi `$, and clearly the density profile is significantly modified even when the area weighted toroidal plasma $`\beta `$ is much greater than one.
## 6 Summary
### 6.1 Steady Jets
Although inclusion of a weak magnetic field (weak in the sense that the flow is highly supermagnetosonic) has little effect on the propagation of adiabatic, overdense jets (e.g. Figures 2 and 3), such fields have strong effects on cooling, overdense jets (e.g. Figures 4, 5, 9 and 10). The difference is caused by the much larger compression of jet and ambient gas that occurs in cooling shocks. This compression amplifies the magnetic field until in some cases the magnetic pressure is comparable to the gas pressure in shocked gas, a situation which never occurs in the adiabatic jet simulations.
Our simulations reveal at least three effects of the magnetic field on steady, cooling jets. Firstly, the field affects the structure and fragmentation of dense sheets of cooled gas (e.g. Figures 4 and 9). This is in part because magnetic pressure can support cooling gas against compression, and so limit the density and width of cooling shells. Alternatively, magnetic stresses can inhibit the radial flow of gas away from the axis, leading to higher densities in some cases. Both of these effects are evident in comparison of models $`\beta _{\mathrm{}}^{Cl}`$, $`\beta _1^{Cl}`$, and $`\beta _{1/4}^{Cl}`$ (Figure 4). The structure of the magnetic field is important to its overall effect, for example, helical magnetic fields appears to produce smoother fluctuations in the cocoon than purely toroidal fields of the same strength (see Figure 9). Much of this may be attributed to the difference the field has on the radial structure of the jet, either through expansion or compression of the jet beam by pressure and hoop stresses. Jets which undergo radial expansion have a much larger cross section near their tip, and drive larger, more fragmented shells into the ambient gas. Jets which are confined or compressed by the magnetic field tend to have narrower and smoother cocoons.
Secondly, the presence of a magnetic field can, in some cases, result in the production of a dense, plug of shocked jet material ahead of the Mach disk (a โnose coneโ). The formation of nose cones has been noted previously in simulations of magnetized extragalactic jets (Clarke et al. 1986). However, we find the formation of nose cones in cooling jets is sensitive to the geometry of the field: only purely toroidal fields which peak near the surface of the jet form nose cones, simply because this configuration maximizes the hoop stresses which confine the nose cone. Moreover, nose cones are likely to unstable in 3D which will limit their relevance to real protostellar jets.
Finally, the presence of magnetic fields affects the stability of the jet beam. We do not observe any evidence for the MHD K-H instability in our simulations, nor do we expect to given that the linear growth rates for such modes are small for the supermagnetosonic jets studied here. Interestingly, we observe axial pressure fluctuations in a few simulations where the jet beam is unstable to current driven pinch modes (Begelman 1998). However, our simulations reveal such modes saturate at low amplitude in the nonlinear regime, so that it is unlikely they will have much relevance to the internal knots observed in most protostellar jets.
### 6.2 Pulsed Jets
Our simulations reveal that weak magnetic fields have a number of effects on time-variable (pulsed) cooling jets. Such pulsing results in the formation of dense knots of shocked jet material in the jet beam. The inclusion of a toroidal magnetic field affects the rate at which such knots grow by inhibiting the radial flow of material out of the knots and into the cocoon. This results in much higher densities in the pulses as the field strength is increased (Figure 13). The increase of the density in the cooling gas in the shocked pulses as the magnetic field strength is increased contradicts the expectations of simple planar shock theory: this result emphasizes the importance of multidimensional effects in MHD flows.
A second effect of a non-uniform magnetic field is the introduction of radial structure into the density and pressure of the pulses, even if the jet is uniform at the nozzle. Jets with strong toroidal fields have radial density profiles in the pulses which are strongly peaked toward the axis. This profile is introduced by radial variations in the magnetic pressure and hoop stresses in the shocked pulses, which confines material near the axis. It is unlikely that real protostellar jets have a uniform (โtop-hatโ) profile: it has been shown that the structure of the cocoon can be affected by the profile of the density and velocity in the jet beam through in hydrodynamical simulations (Suttner et al 1997). Our results reinforce the importance of understanding the radial profile of the outflow to interpret the structures observed.
Finally, even though toroidal fields affect the internal radial structure of the pulses, we do not observe much effect on the structure of the cocoon due to the magnetic field. Radial profiles of various quantities in the pulses (Figure 16) show that the maximum axial velocity of material ejected from the pulses into the cocoon is relatively insensitive to the field strength. Thus, the size of the cocoon is not significantly decreased in the magnetized jets. The radial ejection velocity is small (a few times the sound speed in the unperturbed jet beam), and it would require a delicate force balance to completely eliminate radial outflow.
## 7 Conclusions
We have studied the propagation of both steady and time-variable (โpulsedโ) protostellar jets using numerical MHD simulations. Although we focus our attention on models in which the jet beam contains a purely toroidal magnetic field peaked near its surface, we have also studied the effect of varying both the strength and geometry of the field on the dynamics.
We find that even a weak magnetic field ($`B60\mu `$G) in the jet beam can lead to important effects on the structure and dynamics of steady jets (FRJN-C; CGH; CG). In particular, such fields can alter the density structure and fragmentation of dense shells formed in cooling jets. However, the details of the effects depend sensitively on the geometry of the field. For example, while magnetic pressure can limit compression in such shells, resulting in lower densities and less fragmented shells, radial hoop stresses associated with purely toroidal fields can confine shocked gas towards the axis, resulting in higher densities there. In some cases, hoop stresses lead to the formation of nose cones ahead of the Mach disk. We see no evidence for significant structure induced by the Kelvin-Helmholtz instability (such modes were seen for the parameters and field geometry adopted by CG). We do see evidence for current driven pinch modes but the induced pressure pulses in the jet interior do not appear strong enough to explain the emission knots that are observed in protostellar jets (see also CG). Thus, models invoking jet velocity fluctuation appear to remain the most viable explanation for the knots observed in protostellar outflows. We do not see any evidence for disruption of the jet by pinch modes.
In the case of pulsed cooling jets, we find the primary effect of a toroidal magnetic field is to confine shocked jet material to the axis, preventing it from begin ejected into the cocoon, and leading to higher postshock densities in the pulses in comparison to purely hydrodynamic models. This result is in contrast to the expectation of planar radiative shock models, in which the addition of a magnetic field leads to lower postshock densities. Our results indicate it is important to account for multidimensional effects in the study of magnetized cooling jets. Toroidal confinement also leads to radial variation of quantities in the pulses, even if the density and velocity initially are constant with radius. The radial variation of, e.g. the density, in the pulses is large enough (a factor of 80 in a strongly magnetized jet) that it will likely affect the computation of the resulting emission properties.
While there are many uncertainties in the magnetic field strength and topology implied by the observations, our present results make it clear that MHD models of protostellar jets need to be seriously investigated. This is particularly true since we find that average values of the magnetic field which correspond to a plasma $`\beta `$ much larger than one (and which therefore one would infer to be too weak to be important) can still make a significant difference in the dynamics and physical conditions associated with the jet, and therefore its emission properties as well. It is thus important to understand the asymptotic structure of the field produced by the mechanism which drives the outflow.
We thank E. de Gouveia Dal Pino for comments on an earlier version of the manuscript, and an anonymous referee for suggested improvements. JS acknowledges support from the DOE through grant DFG0398DP00215. PH acknowledges support from the National Science Foundation through grant AST-9318397 and AST-9802955 to the University of Alabama.
## 8 Figure Captions
TABLE 1
Simulation Parameters For Steady Jets
| Simulation | $`M_{j,ms}^{(a)}`$ | $`\overline{M}_{j,ms}^{(b)}`$ | $`M_{a,ms}`$ | $`P_j^{(a)}/P_0`$ | $`r_m/R_j`$ | $`\beta _{\varphi ,m}^{(c)}`$ | $`\beta _z^{(c)}`$ |
| --- | --- | --- | --- | --- | --- | --- | --- |
| $`\beta _{\mathrm{}}^{Ad}`$ | 63.2 | 63.2 | 20 | 1 | | | |
| $`\beta _1^{Ad}`$ | 43.9 | 48.3 | 20 | 2 | 0.9 | 1 | |
| $`\beta _{1/4}^{Ad}`$ | 22.4 | 25.5 | 20 | 8 | 0.9 | 0.25 | |
| $`\beta _{\mathrm{}}^{Cl}`$ | 63.2 | 63.2 | 283 | 1 | | | |
| $`\beta _1^{Cl}`$ | 43.9 | 48.3 | 283 | 2 | 0.9 | 1 | |
| $`\beta _{1/4}^{Cl}`$ | 22.4 | 25.5 | 283 | 8 | 0.9 | 0.25 | |
| $`\beta _{ff}^{Cl}`$ | 14.1 | 30.1 | 283 | 1 | 0.05 | 0.062 | |
| $`\beta _{Rm}^{Cl}`$ | 19.2 | 42.0 | 283 | 11 | 0.2 | 0.25 | |
| $`\beta _{\varphi z}^{Cl}{}_{}{}^{(d)}`$ | 20.9 | 24.5 | 18.3 | 8 | 0.8 | 0.25 | 1 |
<sup>a</sup> on jet axis in first computational zone
<sup>b</sup> linear average
<sup>c</sup> $`\beta =1`$ coresponds to $`B=58.8\mu G`$
<sup>d</sup> 10 zones/R<sub>j</sub>
TABLE 2
Simulation Parameters For Pulsed Jets
| Simulation | $`M_{j,ms}^{(a)}`$ | $`M_{j,ms}^{(b)}`$ | $`M_{j,s}^{(b)}`$ | $`M_{a,ms}`$ | $`P_j^{(a)}/P_0`$ | $`r_m/R_j`$ | $`\beta _{\varphi ,m}^{(c)}`$ | $`\beta _z^{(c)}`$ | $`\beta _\varphi ^{(b)}`$ |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| $`\beta _{\mathrm{}}^P`$ | 63.2 | 63.2 | 63.2 | 283 | 1 | | | | $`\mathrm{}`$ |
| $`\beta _1^P`$ | 43.9 | 50.5 | 68.9 | 283 | 2 | 0.9 | 1 | | 6.7 |
| $`\beta _{1/4}^P`$ | 22.4 | 27.3 | 46.4 | 283 | 8 | 0.9 | 0.25 | | 8.2 |
| $`\beta _{Rm}^P`$ | 33.6 | 57.2 | 60.7 | 283 | 3.5 | 0.2 | 1 | | 16.4 |
| $`\beta _{\varphi z}^P`$ | 34.5 | 38.3 | 64.6 | 18.3 | 2.2 | 0.8 | 1 | 1 | 10.2 |
<sup>a</sup> on jet axis in first computational zone
<sup>b</sup> area weighted average
<sup>c</sup> $`\beta =1`$ coresponds to $`B=58.8\mu G`$ |
warning/0003/quant-ph0003146.html | ar5iv | text | # Untitled Document
First LANL draft, March 30, 2000
Does Quantum Nonlocality Exist?
Bellโs Theorem and the
Many-Worlds Interpretation
by
Frank J. Tipler<sup>1</sup> e-mail address: tipler@mailhost.tcs.tulane.edu
Department of Mathematics and Department of Physics
Tulane University
New Orleans, Louisiana 70118 USA
Abstract
Quantum nonlocality may be an artifact of the assumption that observers obey the laws of classical mechanics, while observed systems obey quantum mechanics. I show that, at least in the case of Bellโs Theorem, locality is restored if observed and observer are both assumed to obey quantum mechanics, as in the Many-Worlds Interpretation. Using the MWI, I shall show that the apparently โnon-localโ expectation value for the product of the spins of two widely separated particles โ the โquantumโ part of Bellโs Theorem โ is really due to a series of three purely local measurements. Thus, experiments confirming โnonlocalityโ are actually confirming the MWI.
PACS numbers: 03.67.Hk, 42.50.-p, 03.65.Bz, 89.70.+c
Nonlocality is the standard example of a quantum mechanical property not present in classical mechanics. Many papers are published each year (in 1997, four in PRL alone ; in 1998, six in PRL alone ; in 1999, eight in PRL alone ; and three in Nature ) on the subject ofโnonlocality,โ many (e.g., the papers just cited) showing truly awesome ingenuity. The phenomenon of nonlocality was first discussed in the EPR Experiment . We have two spin 1/2 particles, and the two-particle system is in the rotationally invariant singlet state with zero total spin angular momentum. Thus, if we decide to measure the particle spins in the up-down direction, we would write the wave function of such a state as
$$|\mathrm{\Psi }>=\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}$$
$`(1)`$
where the direction of the arrow denotes the direction of spin, and the subscript denotes the particle. If we decide to measure the particle spins in the left-right direction, the wave function would be written in a left-right basis as
$$|\mathrm{\Psi }>=\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}$$
$`(2)`$
Nonlocality arises if and only if we assume that the measurement of the spin of a particle โcollapses the wave functionโ from the linear superposition to either $`|>_1|>_2`$ or $`|>_1|>_2`$ in $`(1)`$. If such a collapse occurs, then measuring the spin of particle one would fix the spin of particle two. The spin of particle two would be fixed instantaneously, even if the particles had been allowed to separate to large distances. If at the location of particle one, we make a last minute decision to measure the spin of particle one in the left-right direction rather than the up-down direction, then instantaneously the spin of particle two would be fixed in the opposite direction as particle one โ if we assume that $`(2)`$ collapses at the instant we measure the spin of particle one. The mystery of quantum nonlocality lies in trying to understand how particle two changes โ instantaneously โ in response to what has happened in the location of particle one.
There is no mystery. There is no quantum nonlocality. Particle two doesnโt know what has happened to particle one when its spin is measured. State transitions are nice and local in quantum mechanics. These statements are true because quantum mechanics tells us that the wave function does not collapse when the state of a system is measured. In particular, nonlocality disappears when the Many-Worlds Interpretation (MWI) is adopted. The MWI dispels the mysteries of quantum mechanics. D.N. Page has previously shown how the EPR reality criterion is completely fulfilled by the MWI. I shall extend his analysis, and show how the โquantumโ part of Bellโs Theorem , namely the expectation valve for the product of the spins of the two widely separated electrons, a quantity generally believed to be essentially non-local, actually arises from a series of local measurements.
To see how nonlocality disappears, let us analyze the measure of the spins of the two particles from the Many-Worlds perspective. Let $`M_i(\mathrm{})`$ denote the initial state of the device which measures the spin of the $`i`$th particle. The ellipsis will denote a measurement not yet having been performed. We can for simplicity assume that the apparatus is 100% efficient and that the measurement doesnโt change the spin being measured (putting in a more realistic efficiency and taking into account the fact that measurement may affect the spin slightly would complicate the notation but the conclusions would be unchanged). That is, if each particle happens to be in an eigenstate of spin, a measurement of the $`i`$th particle changes the measuring device โ but not the spin of the particle โ as follows:
$$๐ฐ_1M_1(\mathrm{})|>_1=M_1()|>_1,๐ฐ_1M_1(\mathrm{})|>_1=M_1()|>_1$$
$`(3)`$
$$๐ฐ_2M_2(\mathrm{})|>_2=M_2()|>_2,๐ฐ_2M_2(\mathrm{})|>_2=M_2()|>_2$$
$`(4)`$
where $`๐ฐ_i`$ are linear operators which generates the change of state in the measurement apparatus, corresponding to the measurement.
In particular, if particle 1 is in an eigenstate of spin up, and particle 2 is in an eigenstate of spin down, then the effect of the $`๐ฐ_i`$โs together is
$$๐ฐ_2๐ฐ_1M_1(\mathrm{})M_2(\mathrm{})|>_1|>_2=M_1()M_2()|>_1|>_2$$
$`(5)`$
even if particles 1 and 2 are light years apart when their spin orientations are measured. Similarly, the result of measuring the ith particle in the eigenstate of spin left would be $`๐ฐ_iM_i(\mathrm{})|>_i=M_i()|>_i`$, and for an eigenstate of spin right $`๐ฐ_iM_i(\mathrm{})|>_i=M_i()|>_i`$, which will generate equations for spins left and right analogous to eqs. $`(3)(5)`$.
Now consider the effect of a measurement on the two particle system in the Bohm state, that is, with total spin zero. This state is $`(1)`$ or $`(2)`$ with respect to an up/down or left/right basis respectively. The result is completely determined by linearity and the assumed correct measurements on single electrons in eigenstates. For example, the effect of measurements in which both observers happen to choose to measure with respect to the up/down basis is
$$๐ฐ_2๐ฐ_1M_2(\mathrm{})M_1(\mathrm{})\left[\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}\right]=$$
$$๐ฐ_2M_2(\mathrm{})\left[\frac{M_1()|>_1|>_2}{\sqrt{2}}\frac{M_1()|>_1|>_2}{\sqrt{2}}\right]$$
$$=\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}$$
$`(6)`$
It may appear from eqn. (6) that it is the first measurement to be carried out that determines the split into the two worlds represented by two terms in $`(6)`$. This is false. In fact, if the measurements are carried out at spacetime events which are spacelike separated, then there is no Lorentz invariant way of determining which measurement was carried out first. At spacelike separation, the measuring operators $`๐ฐ_1`$ and $`๐ฐ_2`$ commute, and so we can equally well perform the measurement of the spins of the electrons in reverse order and obtain the same splits:
$$๐ฐ_1๐ฐ_2M_1(\mathrm{})M_2(\mathrm{})\left[\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}\right]=$$
$$๐ฐ_1M_1(\mathrm{})\left[\frac{M_2()|>_1|>_2}{\sqrt{2}}\frac{M_2()|>_1|>_2}{\sqrt{2}}\right]$$
$$=\frac{M_1()M_2()|>_1|>_2}{\sqrt{2}}\frac{M_1()M_2()|>_1|>_2}{\sqrt{2}}$$
$`(7)`$
the last line of which is the same as that of (6), (except for the order of states, which is irrelevant).
The effect of measurements in which both observers happen to choose to measure with respect to the left/right basis is
$$๐ฐ_2๐ฐ_1M_2(\mathrm{})M_1(\mathrm{})\left[\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}\right]=$$
$$๐ฐ_2M_2(\mathrm{})\left[\frac{M_1()|>_1|>_2}{\sqrt{2}}\frac{M_1()|>_1|>_2}{\sqrt{2}}\right]$$
$$=\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}$$
$`(8)`$
A comparison of (6)/(7) with (8) shows that if two spacelike-separated observers fortuitously happen to measure the spins of the two particles in the same direction โ whatever this same direction happens to be โ both observers will split into two distinct worlds, and in each world the observers will measure opposite spin projections for the electrons. But at each event of observation, both of the two possible outcomes of the measurement will be obtained. Locality is preserved, because indeed both outcomes are obtained in total independence of the outcomes of the other measurement. The linearity of the operators $`๐ฐ_1`$ and $`๐ฐ_2`$ forces the perfect anti-correlation of the spins of the particles in each world. Since the singlet state is rotationally invariant, the same result would be obtained whatever direction the observers happened to choose to measure the spins.
In the EPR experiment, there is actually a third measurement: the comparison of the two observations made by the spatially separated observers. In fact, the relative directions of the two spin measurements have no meaning without this third measurement. Once again, it is easily seen that initialization of this third measurement by the two previous measurements, plus linearity implies that this third measurement will confirm the split into two worlds. In the usual analysis, this third measurement is not considered a quantum measurement at all, because the first measurements are considered as transferring the data from the quantum to the classical regime. But in the MWI, there is no classical regime; the comparison of the data in two macroscopic devices is just a much a quantum interaction as the original setting up of the singlet state. Furthermore, this ignored third measurement is actually of crucial importance: it is performed after information about the orientation of the second device has been carried back to the first device (at a speed less than light!). The orientation is coded with correlations of the spins of both electrons, and these correlations (and the linearity of all operators) will force the third measurement to respect the original split. These correlations have not been lost, for no measurement reduces the wave function: the minus sign between the two worlds is present in all eqns. $`(1)`$$`(8)`$.
To see explicitly how this third measurement works, represent the state of the comparison apparatus by $`M_c[(\mathrm{})_1(\mathrm{})_2]`$, where the first entry measures the record of the apparatus measuring the first particle, and the second entry measures the record of the apparatus measuring the second particle. Thus, the third measurement acting on eigenstates of the spin-measurement devices transforms the comparison apparatus as follows:
$$๐ฐ_cM_c[(\mathrm{})_1(\mathrm{})_2]M_1()=M_c[()_1(\mathrm{})_2]M_1()$$
$$๐ฐ_cM_c[(\mathrm{})_1(\mathrm{})_2]M_1()=M_c[()_1(\mathrm{})_2]M_1()$$
$$๐ฐ_cM_c[(\mathrm{})_1(\mathrm{})_2]M_2()=M_c[(\mathrm{})_1()_2]M_2()$$
$$๐ฐ_cM_c[(\mathrm{})_1(\mathrm{})_2]M_2()=M_c[(\mathrm{})_1()_2]M_2()$$
where for simplicity I have assumed the spins will be measured in the up or down direction. Then for the state $`(1)`$, the totality of the three measurements together โ the two measurements of the particle spins followed by the comparison measurement โ is
$$๐ฐ_c๐ฐ_2๐ฐ_1M_c[(\mathrm{})_1(\mathrm{})_2]M_2(\mathrm{})M_1(\mathrm{})\left[\frac{|>_1|>_2|>_1|>_2}{\sqrt{2}}\right]=$$
$$=M_c[()_1()_2]\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}M_c[()_1()_2]\frac{M_2()M_1()|>_1|>_2}{\sqrt{2}}$$
Heretofore I have assumed that the two observers have chosen to measure the spins in the same direction. For observers who make the decision of which direction to measure the spin in the instant before the measurement, most of the time the two directions will not be the same. The experiment could be carried out by throwing away all observations except those in which the chosen directions happened to agree within a predetermined tolerance. But this would waste most of the data. The Aspect-Clauser-Freedman Experiment is designed to use more of the data by testing Bellโs Inequality for the expectation value of the product of the spins of the two electrons with the spin of one electron being measured in direction $`\widehat{๐ง}_1`$, and the spin of the other in direction $`\widehat{๐ง}_2`$. If the spins are measured in units of $`\mathrm{}/2`$, the standard QM expectation value for the product is
$$<\mathrm{\Psi }|(\widehat{๐ง}_1\sigma _1)(\widehat{๐ง}_2\sigma _2)|\mathrm{\Psi }>=\widehat{๐ง}_1\widehat{๐ง}_2$$
$`(9)`$
where $`|\mathrm{\Psi }>`$ is the singlet state (1)/(2). In particular, $`\widehat{๐ง}_1=\widehat{๐ง}_2`$ is the assumed set-up of the previous discussion. Since the MWI shows that local measurements in this case always gives +1 for one electron and -1 for the other, the product of the two is always -1 in all worlds, and thus the expectation value for the product is -1, in complete agreement with (9).
To show how (9) comes about by local measurements splitting the universe into distinct worlds, I follow and write the singlet state (1)/(2) with respect to some basis in the $`\widehat{๐ง}_1`$ direction as
$$|\mathrm{\Psi }>=(1/\sqrt{2})(|\widehat{๐ง}_1,>_1|\widehat{๐ง}_1,>_2|\widehat{๐ง}_1,>_1|\widehat{๐ง}_1,>_2)$$
$`(10)`$
Let another direction $`\widehat{๐ง}_2`$ be the polar axis, with $`\theta `$ the polar angle of $`\widehat{๐ง}_1`$ relative to $`\widehat{๐ง}_2`$. Without loss of generality, we can choose the other coordinates so that the azimuthal angle of $`\widehat{๐ง}_1`$ is zero. Standard rotation operators for spinor states then give
$$|\widehat{๐ง}_1,>_2=(\mathrm{cos}\theta /2)|\widehat{๐ง}_2,>_2+(\mathrm{sin}\theta /2)|\widehat{๐ง}_2,>_2$$
$$|\widehat{๐ง}_1,>_2=(\mathrm{sin}\theta /2)|\widehat{๐ง}_2,>_2+(\mathrm{cos}\theta /2)|\widehat{๐ง}_2,>_2$$
which yields
$$|\mathrm{\Psi }>=(1/\sqrt{2})[(\mathrm{sin}\theta /2)|\widehat{๐ง}_1,>_1|\widehat{๐ง}_2,>_2+(\mathrm{cos}\theta /2)|\widehat{๐ง}_1,>_1|\widehat{๐ง}_2,>_2$$
$$(\mathrm{cos}\theta /2)|\widehat{๐ง}_1,>_1|\widehat{๐ง}_2,>_2(\mathrm{sin}\theta /2)|\widehat{๐ง}_1,>_1|\widehat{๐ง}_2,>_2]$$
$`(11)`$
In other words, if the two devices measure the spins in arbitrary directions, there will be a split into four worlds, one for each possible permutation of the electron spins. Just as in the case with $`\widehat{๐ง}_1=\widehat{๐ง}_2`$, normalization of the devices on eigenstates plus linearity forces the devices to split into all of these four worlds, which are the only possible worlds, since each observer must measure the electron to have spin $`+1`$ or $`1`$.
The squares of the coefficients in $`(11)`$ are proportional to the number of worlds wherein each respective possibility occurs, these possibilities being determined by the chosen experimental arrangement. This is most easily seen using Deutschโs MWI derivation of the Born Interpretation (BI). DeWitt and Graham originally deduced the BI using the relative frequency theory of probability , and this derivation is open to the standard objections to the frequency theory . Deutsch instead derives the BI using the Principle of Indifference of the classical/a priori theory of probability . According to the Principle of Indifference, the probability of an event is the number of times the event occurs in a collection of equipossible cases divided by the total number of equipossible cases. Thus, the probability is 1/6 that a single die throw will result in a 5, because there are 6 equipossible sides that could appear, of which the 5 is exactly 1.
Deutsch assumes the Principle of Indifference applies to any experimental arrangement in which the expansion of the wave function $`|\mathrm{\Psi }>`$ in terms of the orthonormal basis vectors of the experiment (the interpretation basis) give equal coefficients for each term in the expansion. For example, both $`(1)`$ and $`(2)`$ are two such expansions, because in both cases the coefficients of each of the two terms is $`1/\sqrt{2}`$. The Principle of Indifference thus says that each of the two possibilities is equally likely in either the experimental arrangement $`(1)`$, or in the interpretation basis $`(2)`$. Equivalently, there are an equal number of worlds corresponding to each term in either $`(1)`$ or $`(2)`$, since in the MWI โequally possibleโ means โequal number of worldsโ (equal relative to a preset experimental arrangement).
Deutsch shows that if the squares of the coefficients in the interpretation basis of an experiment are rational, then a new experimental arrangement can be found in which the coefficients are equal in the new interpretation basis. Applying the Principle of Indifference to this new set of coefficients yields the BI for the coefficients in the original basis. Continuity in the Hilbert space of wave functions yields the BI for irrational coefficients (although it is a presupposition of the MWI that only coefficients with rational squares are allowed since irrational squares would imply an irrational number of worlds). In particular, the percentage of worlds with the value of a given basis vector is given by the square of the coefficient.
The expectation value $`(9)`$ for the product of the spins is just the sum of each outcome, multiplied respectively by probabilities of each of the four possible outcomes:
$$(+1)(+1)P_{}+(+1)(1)P_{}+(1)(+1)P_{}+(1)(1)P_{}$$
$`(12)`$
where $`P_{}`$ is the relative number of worlds in which the first electron is measured spin up, and the second electron spin down, and similarly for the other $`P`$โs. Inserting these relative numbers โ the squares of the coefficients in $`(11)`$ โ into $`(12)`$ gives the expectation value:
$$=\frac{1}{2}\mathrm{sin}^2\theta /2\frac{1}{2}\mathrm{cos}^2\theta /2\frac{1}{2}\mathrm{cos}^2\theta /2+\frac{1}{2}\mathrm{sin}^2\theta /2=\mathrm{cos}\theta =\widehat{๐ง}_1\widehat{๐ง}_2$$
$`(13)`$
which is the quantum expectation value $`(9)`$.
Once again it is essential to keep in mind the third measurement that compares the results of the two measurements of the spins, and by bringing the correlations between the worlds back to the same location, defines the relative orientation of the previous two measurements, and in fact determines whether there is a twofold or a fourfold split. The way the measurement of $`(9)`$ is actually carried out in the Aspect-Clauser-Freedman Experiment is to let $`\theta `$ be random in any single run, and for the results of each fixed $`\theta `$ from a series of runs be placed in separate bins. This separation requires the third measurement, and this local comparison measurement retains the correlations between the spins. The effect of throwing away this correlation information would be equivalent to averaging over all $`\theta `$ in the computation of the expectation value: the result is $`_0^\pi <\mathrm{\Psi }|(\widehat{๐ง}_1\sigma _1)(\widehat{๐ง}_2\sigma _2)|\mathrm{\Psi }>๐\theta =0`$; i.e., the measured spin orientations of the two electrons are completely uncorrelated. This is what we would expect if each measurement of the electron spins is completely local, which in fact they are. There is no quantum nonlocality.
Bellโs results lead one to think otherwise. But Bell made the tacit assumption that each electronโs wave function is reduced by the measurement of its spin. Specifically, he assumed that the first electronโs spin was determined by the measurement direction $`\widehat{๐ง}_1`$ and the value some local hidden variable parameters $`\lambda _1`$: the first electronโs spin is given by a function $`A(\widehat{๐ง}_1,\lambda _1)`$. The second electronโs spin is given by an analogous function $`B(\widehat{๐ง}_2,\lambda _2)`$, and so the hidden variable expectation value for the product of the spins would not be $`(13)`$ but instead
$$\rho (\lambda _1,\lambda _2)A(\widehat{๐ง}_1,\lambda _1)B(\widehat{๐ง}_2,\lambda _2)๐\lambda _1๐\lambda _2$$
$`(14)`$
where $`\rho (\lambda _1,\lambda _2)`$ is the joint probability distribution for the hidden variables. By comparing a triple set of directions $`(\widehat{๐ง}_1,\widehat{๐ง}_2,\widehat{๐ง}_3)`$, Bell derived the inequality $`|P(\widehat{๐ง}_1,\widehat{๐ง}_2)P(\widehat{๐ง}_1,\widehat{๐ง}_3)|1+P(\widehat{๐ง}_2,\widehat{๐ง}_3)`$, which for certain choices of the triple, is inconsistent with the quantum mechanical $`(9)`$; i.e., $`\widehat{๐ง}_1=\widehat{๐ง}_2\widehat{๐ง}_3/|\widehat{๐ง}_2\widehat{๐ง}_3|`$ yields $`\sqrt{2}1`$ if we assume the MW result $`(9)/(13)`$, which is the quantum part of Bellโs Theorem.
But $`(14)`$ assumes that the spin of each particle is a function of $`\widehat{๐ง}_i`$ and $`\lambda _i`$; that is, it assumes the spin at a location is single-valued. This is explicitly denied by the MWI, as one can see by letting $`\lambda _i`$ be the spatial coordinates of the ith electron. Bellโs analysis tacitly assumes that the macroscopic world is a single-valued world like classical mechanics.
The automatic elimination of action at a distance by the MWI is a powerful argument for the validity of the MWI, for assuming that both single electrons and many-atom measuring devices are described by multivalued quantum states.
I thank R. Chiao, D. Deutsch, B. DeWitt, and D.N. Page for helpful discussions, and M. Millis for inviting me to speak at a NASA conference where questions of nonlocality were discussed.
References
M. Horodecki, et al, Phys. Rev. Lett. 78, 574 (1997); A. Zeilinger, et al, PRL 78, 3031 (1997); E. Hagley, et al, PRL 79, 1 (1997); D. Boschi et al, PRL 79, 2755 (1997); B. Yurke et al, PRL 79, 4941 (1997).
M. Lewenstein, et al, Phys. Rev. Lett. 80, 2261 (1998), quant-ph/9707043; A. Gilchrist, et al, PRL 80, 3169 (1998); J.-W. Pan, et al, PRL 80, 3891 (1998); F. De Martini, PRL 81, 2842 (1998), quant-ph/9710013; W. Tittle et al, PRL 81, 3563 (1998), A. Zeilinger et al, PRL 81, 5039 (1998), quant-ph/9810080.
D. Bouwmeester et al, PRL 82, 1345 (1999), quant-ph/9810035; K. Banaszek et al, PRL 82, 2009 (1999), quant-ph/9910117; A. Bramon et al, PRL 83, 1 (1999), hep-ph/9811406; N. Linden et al, PRL 83, 243 (1999), quant-ph/9902022; R. Polkinghorne et al, PRL 83, 2095 (1999), quant-ph/9906066; S. Aerts et al, PRL 83, 2872 (1999), quant-ph/9912064; A. White et al, PRL 83, 3103 (1999), quant-ph/9908081; D.A. Meyer, PRL 83, 3751 (1999).
D. Bouwmeester et al, Nature 390, 575 (1997), 403, 515 (2000); C.A. Sackett et al, Nature 404, (2000).
A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 47, 777 (1935).
H. Everett, Rev. Mod. Phys. 29, 454 (1957).
B.S. DeWitt and N. Graham, ed., The Many-Worlds Interpretation of Quantum Mechanics (Princeton U.P., Princeton, 1973)
S. Goldstein and D.N. Page, PRL 74, 3715 (1995).
D.N. Page, Phys. Lett. 91A, 57 (1982).
J.S. Bell, Physics 1, 195 (1964).
S.J. Freedman and J.F. Clauser, PRL 28, 938 (1972); A. Aspect et al, PRL 47, 1804 (1982).
D.M. Greenberger et al, Am. J. Phys. 58, 1131 (1990).
D. Deutsch, Proc. Roy. Soc. London A (original preprint 1989), quant-ph/9906015. (B. DeWittโs outline (ref. below), is shorter and may be easier to understand).
T.L. Fine, Theories of Probability (Academic Press, New York, 1973); R. Weatherford, Philosophical Foundations of Probability Theory (Routledge, London, 1982).
M. Jammer, The Philosophy of Quantum Mechanics (Wiley, NY, 1974).
B.S. DeWitt, Int. J. Modern Phys. A13, 1881โ1916 (1998). |
warning/0003/physics0003088.html | ar5iv | text | # Radiative recombination of bare Bi83+: Experiment versus theory
## I Introduction
Recombination between electrons and highly charged ions plays an important role in different areas of modern physics. The basic two- and three-body recombination processes are of very fundamental nature and thus provide an excellent testing ground for collision theory and atomic structure calculations. Cross sections and rate coefficients of these processes are needed for the understanding of astrophysical and fusion plasmas and also provide useful information for applications in accelerator physics . In particular, beam losses in ion storage rings by electron-ion recombination during electron cooling can post harsh limits to the handling and the availability of ions for further experiments.
During the last decade electron-ion recombination has been extensively investigated. Merged-beams experiments using storage ring coolers and single-pass electron targets at accelerators have provided excellent conditions for a new generation of recombination measurements. In these experiments an incident ion beam is merged with a cold beam of electrons over a distance of typically 50 to 250 cm depending on the specific electron beam device. By choosing the appropriate accelerator facility, ions of most elements in all possible charge states can be investigated nowadays.
Free electrons can be captured by ions via several different mechanisms. The main recombination channel for a bare ion is radiative recombination (RR)
$$e+A^{q+}A^{(q1)+}+h\nu .$$
(1)
RR is the direct capture of a free electron by an ion A<sup>q+</sup> where the excess energy and momentum are carried away by a photon. After the capture, which is inverse to photoionization, the electron can be in an excited state and there will be further radiative transitions within the ion until the electron has reached the lowest accessible energy level. RR is a non-resonant process with a diverging cross section at zero center-of-mass (c.m.) energy which continuously decreases towards higher c.m. energies.
Another recombination mechanism possible for a bare ion is three-body recombination (TBR)
$$A^{q+}+e+eA^{(q1)+}+e^{}$$
(2)
where the excess energy and momentum are carried away by a second electron. This process is important at high electron densities and very low center-of-mass energies between electrons and ions.
The pioneering experiment on radiative recombination of bare ions was performed by Andersen et al. in 1990. Absolute rate coefficients were measured for C<sup>6+</sup> with a merged-beam technique finding a reasonably good agreement between experiment and theory in the investigated energy range from E<sub>rel</sub>= 0 to 1 eV where E<sub>rel</sub> is the relative energy between the ensembles of electrons and ions in the interaction region (for the definition of E<sub>rel</sub> see section III).
In a number of consecutive measurements different bare ions (D<sup>+</sup>,He<sup>2+</sup>,C<sup>6+</sup>,N<sup>7+</sup>,Ne<sup>10+</sup>,Si<sup>14+</sup>,Ar<sup>18+</sup>) have been investigated at several facilities. The measured rate coefficients have been in accordance with a theoretical approach to RR by Bethe and Salpeter for relative energies $`E_{rel}`$ 0.01 eV. The experiments were limited by counting statistics to an energy range of only a few eV. Recently an experiment with Cl<sup>17+</sup>-ions revealed excellent agreement between the measured rate and RR theory for relative energies from 0.01 eV up to 47 eV . However, in all of these measurements strong deviations of the experimental findings from the theoretical predictions were found at very low electron-ion relative energies ($`E_{rel}0.01`$ meV) . The measured rate coefficient typically shows an additional increase towards lower energies resulting in a so-called rate enhancement factor $`ฯต=\alpha _{exp}/\alpha _{RR}`$ at $`E_{rel}`$ = 0 eV of 1.6 (He<sup>2+</sup>) to 10 (Ar<sup>18+</sup>) for bare ions (no enhancement observed with D<sup>+</sup>) and up to 365 for a multicharged multi-electron system like Au<sup>25+</sup> . It should be noted that in contrast to the RR cross section which diverges at $`E_{rel}`$ = 0 eV the measured rate coefficient obtains a finite value due to the experimental electron and ion velocity spreads (see section 3).
After the first observation of the so called enhancement phenomenon in an experiment with U<sup>28+</sup>-ions at the GSI in Darmstadt in 1989 and a later experimental confirmation of that same result this effect has been observed over and over again in many different experiments at different facilities. Whereas the very high rate enhancement factors of multicharged complex ions like Au<sup>25+</sup>, Au<sup>50+</sup>, Pb<sup>53+</sup> and U<sup>28+</sup> could be partly traced back to the presence of additional recombination channels such as dielectronic recombination (DR) or polarization recombination (PIR) the origin of the observed discrepancies between experiment and theory for bare ions is still unknown. Theoretical calculations applying $`MolecularDynamics`$ (MD) computer simulations have been performed by Spreiter et al. . Taking into account the experimental conditions at the heavy ion storage ring TSR in Heidelberg they found an increase of the local electron density around the ion. This could possibly lead to a higher number of recombination processes but to date the recombination of the electrons with the ion is not included in their theoretical description.
For the clarification of this phenomenon dependences of the enhancement on external experimental parameters have been studied at different storage rings. Variations of the electron density within a total range from about 10<sup>6</sup> cm<sup>-3</sup> up to almost 10<sup>10</sup> cm<sup>-3</sup> showed little effect on the enhancement. The dependence of the total recombination rate followed the $`T_{}^{1/2}`$ dependence on the mean transverse electron energy spread $`T_{}`$ as expected for RR alone. A systematic study of the ion charge-state dependence of the excess rate coefficient for a number of bare ions yielded roughly a Z<sup>2.8</sup> scaling for atomic numbers 1 $``$ Z $``$ 14. The first external parameter observed to influence the enhancement has been the magnetic field strength in the interaction region of electrons and ions . The experimental results show a clear increase of the recombination rate maximum of Au<sup>25+</sup> ions at $`E_{rel}=0`$ with increasing magnetic field strength. Recently this dependence has also been observed in experiments with lithium-like F<sup>6+</sup> and bare C<sup>6+</sup>-ions at the TSR heavy ion storage ring in Heidelberg .
In the present measurement, which was carried out at the Experimental Storage Ring of the GSI in Darmstadt, we extended the range of investigated bare ions to the high-Z ion Bi<sup>83+</sup>. It was the first experiment of this kind with such a heavy bare ion providing information on RR in addition to studies of the radiative electron capture in ion-atom collisions using x-ray spectroscopy . Beside a comparison of measured absolute rate coefficients with calculated RR rates for relative energies from $`E_{rel}`$ = 0 eV to 125 eV dependences of the recombination rate on the electron density and the magnetic guiding field of the electron beam in the cooler have been investigated.
The present paper is organized as follows. After a short description of the theoretical approaches, the experimental setup and the evaluation of recombination rates are described. The experimental results are presented and compared with RR calculations.
## II Theory
In order to describe RR theoretically Kramers developed a semi-classical theory already in 1923 . A full quantum mechanical treatment was performed by Stobbe seven years later . In 1957, Bethe and Salpeter derived an approximate formula for the RR cross section that is identical to Kramers result
$$\sigma _{RR}(n,E_{cm})=(2.1\times 10^{22}cm^2)\frac{E_0^2}{nE_{cm}(E_0+n^2E_{cm}).}$$
(3)
The capture of an electron by a bare ion produces a hydrogenic state with principal quantum number $`n`$. In this case E<sub>0</sub> is the binding energy of the ground state electron in the hydrogenic ion or atom and E<sub>cm</sub> is the kinetic energy in the electron-ion center-of-mass frame. The total cross section for this process is obtained by summing up the contributions of all accessible Rydberg states:
$$\sigma _{RR}(E_{cm})=\underset{n=1}{\overset{n_{max}}{}}\sigma _{RR}(n,E_{cm})$$
(4)
where $`n_{max}`$ is the maximum principal quantum number that can contribute. This number is generally limited by experimental conditions. The approach of Bethe and Salpeter clearly shows the typical features of RR cross sections: the divergence at zero electron energy and a monotonic decrease for increasing electron energy. However, as a semi-classical approximation Eq. 3 is only valid in the limit of high quantum numbers and low electron energies, i.e. for $`n1`$ and $`E_{cm}Z^2/n^2Ry`$. Since the quantum mechanical treatment of Stobbe involves the rather tedious evaluation of hydrogenic dipole matrix elements one often applies correction factors $`G_n(E_{cm})`$, the so called Gaunt factors, to Eq. 3 to account for deviations from the correct quantum result at low $`n`$ and high $`E_{cm}`$. The use of Gaunt factors is convenient because they are either tabulated or given in an easy parametrization . We here apply tabulated values $`k_n=G_n(0)`$ using
$`\sigma _{RR}(E_{cm})`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{n_{max}}{}}}(2.1\times 10^{22}cm^2)k_n\times `$ (6)
$`{\displaystyle \frac{(Z^2Ry)^2}{nE_{cm}(Z^2Ry+n^2E_{cm})}}.`$
where $`Z`$ is the nuclear charge of the ion and $`Ry`$ = 13.6 eV is the ground-state energy of the hydrogen atom. Eq. 6 is exact for bare ions at zero relative energy and does not deviate by more than about 5% from the quantum mechanically correct hydrogenic result at the highest energies considered in this paper.
For high-Z ions and high relative energies the non-relativistic dipole approximation presented above is not valid. In general an exact relativistic calculation with the inclusion of higher multipoles is required . Nevertheless experiments investigating the Radiative Electron Capture (REC) into the K-shell of high-Z ions have shown that the experimental results are in very good overall agreement with the non-relativistic dipole approximation as long as $`E_{rel}`$ is below the electron rest energy $`m_ec^2`$. Because of this observed accordance Eq. 6 is used for the calculations presented in this paper.
## III Basic relations
In the experiment cross sections at very low center-of-mass energies $`E_{cm}`$ are not accessible due to the finite velocity spread of the colliding particles. In the present work the lowest limit of $`E_{cm}`$ is at about 1$`\times `$10<sup>-5</sup> eV. The measured quantity is a rate coefficient $`\alpha `$ which theoretically results from a convolution of the cross section with the velocity distribution function $`f(\stackrel{}{v})`$
$$\alpha (v_{rel})=\sigma (v)vf(v_{rel},\stackrel{}{v})d^3v.$$
(7)
In our experiment the ion velocity distribution is negligibly narrow compared to that of the electrons due to the cooling of the ion beam to a relative momentum spread below 10<sup>-4</sup> and due to the large mass difference between electrons and ions. Therefore the distribution function $`f(\stackrel{}{v})`$ is dominated by the electron velocity distribution. Considering the axial symmetry of the merged beams experiment two velocity coordinates are sufficient to describe the distribution: $`v_{}`$ the velocity component in beam direction and $`v_{}`$ the velocity component perpendicular to the beam. The appropriate velocity (or energy) spreads are characterized by two corresponding temperatures $`T_{}`$ and $`T_{}`$. Due to the acceleration of the electrons these temperatures are quite different ($`T_{}`$ $``$ $`T_{}`$) resulting in a highly anisotropic velocity distribution $`f(\stackrel{}{v})`$ which is therefore often called โflattenedโ. Its mathematical form is given by
$`f(v_{rel},\stackrel{}{v})`$ $`=`$ $`{\displaystyle \frac{m_e}{2\pi kT_{}}}\mathrm{exp}({\displaystyle \frac{m_ev_{}^2}{2kT_{}}})\times `$ (9)
$`\sqrt{{\displaystyle \frac{m_e}{2\pi kT_{||}}}}\mathrm{exp}({\displaystyle \frac{m_e(v_{||}v_{rel})^2}{2kT_{||}}})`$
with $`v_{rel}`$ being the average longitudinal center-of-mass velocity
$$v_{rel}=v_{e,||}v_{i,||}/(1v_{i,||}v_{e,||}/c^2).$$
(10)
where $`v_{e,}`$ and $`v_{i,}`$ are the longitudinal velocity components of the electron and ion beam in the laboratory frame, respectively. The relative energy of electrons and ions is
$$E_{rel}=(\gamma _{rel}1)m_ec^2$$
(11)
with
$$\gamma _{rel}=\left[1\left(v_{rel}/c\right)^2\right]^{1/2}.$$
(12)
Under cooling conditions $`v_{rel}`$ is zero and, hence, also E<sub>rel</sub> is zero then. For comparisons of experimental results with theory the cross sections resulting from Eq. 6 have to be convoluted with the experimental velocity distribution according to Eq. 7 using the experimental temperatures $`T_{||}`$ and $`T_{}`$.
## IV Experiment
The measurements have been performed at the Experimental Storage Ring (ESR) of the Gesellschaft fรผr Schwerionenforschung (GSI) in Darmstadt . 295.3 MeV/u Bi<sup>83+</sup>-ions supplied by the GSI linear accelerator UNILAC and the heavy ion synchrotron SIS were injected into the ESR. Only one shot of ions was sufficient to provide an ion current of typically 400-800 $`\mu `$A at the beginning of a measurement. In the storage ring the circulating Bi<sup>83+</sup> ions were merged with the magnetically guided electron beam of the electron cooler with an electron energy of 162 kV (Fig. 1). Before starting a measurement, the ion beam was cooled for several seconds until the beam profiles reached their equilibrium widths. During the measurement the electron energy was stepped through a preset range of values different from the cooling energy thus introducing non-zero mean relative velocities $`v_{rel}`$ between ions and electrons. After each voltage step a cooling interval was inserted. The whole scheme of energy scan measurements was realized by applying voltages from -5 kV to 5 kV to the two drift tubes surrounding the electron and the ion beam in the interaction region. The voltages were supplied by a system of sixteen individual power supplies controlled by very fast high-voltage switches. This instrument has specially been constructed for the recombination experiments . Only 2 ms are needed by this device to switch to and set a certain voltage with a relative precision of 10<sup>-4</sup>. The repetition rate of voltage settings with this precision is limited to 40 per second.
Recombined Bi<sup>82+</sup> ions were counted as a function of the electron energy on a scintillator detector located behind the first dipole magnet downstream of the electron cooler. The dipole magnet bends the circulating Bi<sup>83+</sup> ion beam onto a closed orbit and separates the recombined Bi<sup>82+</sup> ions from that orbit. In between two measurement steps of 40 ms duration, the electron energy was always set to the cooling energy for 20 ms in order to maintain good ion beam quality. The experimental data stream was continuously collected and stored after each ms. The time-resolved measurements allowed us to observe and eliminate drag force effects from the data in a detailed off-line analysis. Such effects are a result of the cooling force exerted by the electrons on the ion beam which can lead to a time-dependent shift of the ion velocity towards the electron velocity. The friction forces are particularly effective at relative energies close to zero.
The kinetic energy $`E_e`$ of the electrons is defined by the cathode voltage $`U_{gun}`$, the drift tube voltage $`U_{drift}`$ and the space charge potential $`U_{sp}`$ in the interaction region. Assuming coaxial beams it is calculated as
$`E_e`$ $`=`$ $`eU_{gun}+eU_{drift}eU_{sp}`$ (13)
$`=`$ $`eU_{gun}+eU_{drift}{\displaystyle \frac{I_er_em_ec^2}{ev_e}}\left[1+2\mathrm{ln}(b/a)\right],`$ (14)
where $`r_e`$ is the classical electron radius. The quantities $`b`$ = 10 cm and $`a`$ = 2.54 cm are the radii of the drift tube and the electron beam, respectively. The ion beam diameter is only of the order of a millimeter and, hence, the electron energy distribution probed by the ions is rather flat across the ion beam. In the present experiment we performed measurements with different electron currents $`I_e`$ which produced space charge potentials ranging from 17.2 V to 51.7 V.
The space charge corrected electron energy $`E_e`$ and the ion energy $`E_i`$ are used to calculate the relative energy of electrons and ions in the center-of-mass frame. A relativistic transformation yields
$`E_{rel}`$ $`=`$ $`\sqrt{(m_ic^2+m_ec^2)^2+}`$ (16)
$`\overline{2\left(E_iE_e+E_im_ec^2+E_em_ic^2A\right)}B,`$
$`A`$ $`=`$ $`\sqrt{E_i(E_i+2m_ic^2)}\sqrt{E_e(E_e+2m_ec^2)}\mathrm{cos}(\varphi )`$ (17)
$`B`$ $`=`$ $`(m_ic^2+m_ec^2)`$ (18)
where $`\varphi `$ is the angle between the electron and the ion beam directions. According to Eq. 16 the minimum relative energy $`E_{rel}`$ = 0 eV cannot be reached if an angle $`\varphi `$ 0 is present. Therefore the alignment of the beams was optimized before the recombination experiments in order to achieve $`\varphi `$ = 0 mrad with an uncertainty of 0.1 mrad.
The counting rate measured at the scanning energy $`E_{meas}`$ is given by
$$R(E_{meas})=\frac{\alpha (E_{meas})\eta Ln_e(E_{meas})N_i}{C\gamma ^2}+R_{back}$$
(20)
with $`\alpha `$ denoting the electron-ion recombination rate coefficient, $`\eta `$ the detection efficiency of the scintillator detector which is very close to unity, $`L=2.5`$ m the nominal length of the interaction zone, $`n_e`$(E) the electron density at energy $`E`$, $`N_\mathrm{i}`$ the number of stored ions, $`C=108.36`$ m the ring circumference and $`\gamma `$ the relativistic Lorentz factor for the transformation between the c. m. and the laboratory frames. $`R_{back}`$ denotes the measured background rate due to collisions with residual gas molecules. In order to extract an absolute rate coefficient from the experimental data the background has to be subtracted by taking into account the counting rate
$$R(E_{ref})=\frac{\alpha (E_{ref})\eta Ln_e(E_{ref})N_i}{C\gamma ^2}+R_{back},$$
(21)
at a reference energy $`E_{ref}`$. Combining Eqs. 20 and 21 $`\alpha `$ at $`E_{meas}`$ has to be calculated from
$`\alpha (E_{meas})`$ $`=`$ $`{\displaystyle \frac{(R(E_{meas})R(E_{ref}))\gamma ^2}{\eta Ln_\mathrm{e}(E_{meas})N_\mathrm{i}/C}}`$ (23)
$`+\alpha (E_{ref}){\displaystyle \frac{n_\mathrm{e}(E_{ref})}{n_\mathrm{e}(E_{meas})}}.`$
Because of RR one always has a non-zero recombination rate coefficient $`\alpha (E_{ref})`$ at the reference point. Usually $`E_{ref}`$ is chosen such that $`\alpha (E_{ref})`$ practically equals zero, but in general one has to re-add the rate which has been neglected by subtracting $`R(E_{ref}`$) from $`R(E_{meas})`$. In the present experiment the reference rate has been measured at the maximum accessible scan energy $`E_{ref}`$= 125 eV. According to RR theory the rate coefficient at this energy is $`\alpha (E_{ref})`$= 3.6$`\times `$ 10<sup>-10</sup> cm<sup>3</sup> s<sup>-1</sup>. This correction leads to a modification of the measured rate coefficient at $`E_{rel}`$= 0 eV by only 0.2 $`\%`$.
Electron and ion beams are merged and demerged by bending the electron beam in a toroidal magnetic field with a bending radius of 120 cm. An electron beam of 2.54 cm radius is still overlapping the ion beam for 25 cm before and after the straight overlap section of 250 cm. The merging and demerging sections therefore contribute to the measured counting rate. As one can see in the left panel of Fig. 2 where the calculated potential distribution of the drift tubes for an applied voltage of 1 V is plotted against the position inside the cooler (along the central axis of ion and electron beam) the influence of the voltage applied to the drift tubes is restricted to the straight overlap section of the cooler. Thus, the electron energy in the toroidal sections is always the same independent of the drift tube potential. This results in a constant contribution to the measured counting rate which is considered by the background subtraction procedure described above.
The right panel of Fig. 2 shows the distribution of angles $`\varphi `$ between electron trajectories and the ion beam direction along the geometrical cooler axis. The electrons strictly following the magnetic field lines, the non-zero angles result from the measured transverse magnetic guiding field components in this section. The measurement was restricted to a length of 2.26 m (90 % of the straight overlap section) fully including the drift tube area. As one can see $`\varphi `$ increases rapidly at the edges of the measured range. Using steerer magnets the electron beam can be shifted to a position minimizing the influence of angular deviations. Both the distributions of the electric potential and angle (Fig. 2) can be combined via Eq. 16 into a distribution of relative energies along the ion beam axis. Fig. 3 shows the relative energies along the straight overlap section for different voltages applied to the drift tubes. Obviously the desired relative energies are only realized along a certain fraction of the whole interaction length which, in addition, depends on the energy. The measured rate coefficient at a given relative energy $`E_{meas}`$ always contains contributions from other relative energies according to:
$$\alpha (E_{meas})=\frac{1}{L}_0^L๐l\alpha (E_{rel}(l))$$
(24)
with $`E_{rel}(l)`$ being the relative energy at the position $`l`$ inside the cooler. According to Eq. 24 the correct rate coefficient can be obtained by a deconvolution which is performed iteratively. In a first iteration step the measured rate coefficient $`\alpha (E_{meas})=\alpha ^{(0)}`$ is inserted as $`\alpha (E_{rel}(l))`$ in Eq. 24. Then the difference $`\mathrm{\Delta }\alpha ^{(0)}`$ between the obtained result and the measured rate is subtracted from $`\alpha ^{(0)}`$. In a next step the new $`\alpha ^{(1)}=\alpha ^{(0)}\mathrm{\Delta }\alpha ^{(0)}`$ is likewise inserted in Eq. 24 and the difference $`\mathrm{\Delta }\alpha ^{(1)}`$ is calculated. The procedure is carried on until in a step k the relative difference $`\mathrm{\Delta }\alpha ^{(k)}(E_{meas})/\alpha (E_{meas})`$ is below $`10^3`$ at all measured energies. This is the case after only a few iteration steps.
Although the relative statistical errors of the results presented below amount to less than $`1\%`$ in the rate coefficient maximum, the systematic uncertainty in the absolute recombination rate coefficient has been estimated to be $`\pm 23\%`$.
## V Experimental results and discussion
In Fig. 4 the measured absolute rate coefficient of Bi<sup>83+</sup> with free electrons is plotted versus the relative energy from 0 eV to 125 eV. The spectrum shows the typical shape of an RR rate coefficient with a maximum at $`E_{rel}`$ = 0 eV and a continuous decrease for increasing relative energies.
For the comparison of the measured rates with RR theory one has to know the maximum principal quantum number $`n_{max}`$ and the temperatures $`T_{||}`$ and $`T_{}`$ (see sections II and III). For the temperature $`T_{}`$ characterizing the transverse motion of the electrons one has to assume $`kT_{}`$ = 120 meV which corresponds to the cathode temperature. For the longitudinal electron motion $`kT_{||}`$ = 0.1 meV is inferred from the analysis of resonance shapes in dielectronic recombination (DR) measurements with lithium-like Bi<sup>80+</sup>-ions performed during the same beamtime. $`n_{max}`$ is determined by field ionization in the dipole magnet which separates the parent beam and the recombined ions. A first approximation of this value is the field ionization limit $`n_F`$, which is obtained from
$$n_F=\left(7.3\times 10^{10}\mathrm{Vm}^1\frac{q^3}{F}\right)^{1/4}$$
(25)
, where q is the charge state of the ion and $`F=v_{i,||}B_{}`$ the motional electric field seen by the ions with velocity $`v_{i,||}`$ in the transverse magnetic field $`B_{}`$ of the charge-analyzing magnet. In this context one also has to account for the possibility that high Rydberg states can decay to states below $`n_F`$, provided the ions have some time between the recombination process and the arrival at the ionizing electric field $`F`$. Therefore, a realistic estimate for the cut-off is given by
$$n_{max}=\mathrm{max}(n_\gamma ,n_F),$$
(26)
where $`n_\gamma `$ denotes the maximum principal quantum number of Rydberg states which decay before the recombined ions arrive at the analyzing magnet and thus are saved from field ionization. A crude estimate of $`n_\gamma `$ based on a number of assumptions on the population and the decay of excited states is obtained from a numerical solution of the following equation
$`n_\gamma `$ $`=`$ $`Z^{4/5}\{2.142\times 10^{10}s^1t_F\left[\kappa (n_\gamma )\right]^2`$ (28)
$`[0.04+\mathrm{ln}(n_\gamma )\mathrm{ln}\kappa (n_\gamma )]\}^{1/5},`$
$`\kappa (n_\gamma )`$ $`=`$ $`1.6+0.018n_\gamma .`$ (29)
The time of flight $`t_F`$ of the ions between the recombination act and the arrival at the analyzing magnet influences the survival probability of ions in high Rydberg states. In the derivation of Eq. 28 cascade processes were neglected. They should not be of large influence considering the decay rates in undisturbed hydrogenic systems. A more problematic assumption underlying Eq. 28 is that the fields seen by the ions during their flight time are considered not to change the nl distribution of states as it results from the initial population by radiative recombination. Changes of decay rates by Stark mixing in the fields are not accounted for but probably deserve further attention. Also, instead of a sharp cut-off at $`n_\gamma `$ or $`n_F`$ a distribution of field ionization probabilities around $`n_\gamma `$ or $`n_F`$, respectively, would be more realistic, however, these cannot easily be calculated without further assumptions. In addition, the influences of the experimental conditions providing electric fields via $`\stackrel{}{v}\times \stackrel{}{B}`$ contributions inside the interaction area, the toroidal fields and the dipole correction magnets after the cooler have to be considered.
In the present experiment $`n_F`$ and $`n_\gamma `$ have been calculated to be 116 and 442, respectively. A comparison of both RR rate curves resulting from Eqs. 6 and 7 is shown in Fig. 5. As expected the theoretical rate for $`n_{max}`$ = $`n_\gamma `$ = 442 (dotted line) lies above the curve with $`n_{max}`$ = $`n_F`$ = 116 (solid line). Comparing the experimental and the theoretical rate coefficients the curve calculated for $`n_F`$ = 116 shows a very good agreement with the experimental data for relative energies from $`E_{rel}`$ = 15 meV to 125 eV (Fig. 4). However, one can also obtain a good agreement for $`n_\gamma `$ = 442 if one assumes a higher transverse electron temperature $`kT_{}`$ = 250 meV. Such an interdependency between $`n_{max}`$ and $`kT_{}`$ has also been found in experiments with bare Cl<sup>17+</sup> and C<sup>6+</sup>-ions at the TSR in Heidelberg . Because of the impossibility to accurately obtain both parameters from a fit of the theoretical RR curve to the experimental spectrum and in view of the rather crude estimation of $`n_{max}`$ we deliberately choose $`n_{max}`$ = $`n_F`$ = 116 throughout the rest of this paper. This implies a reasonable choice for the transverse temperature of $`kT_{}`$ = 120 meV, which is in accordance with the cathode temperature. That same temperature is also suggested by the accompanying DR measurements with Bi<sup>80+</sup>-ions. In this discussion the longitudinal temperature $`T_{}`$ does not play a role since it has very little influence on the RR rate coefficient.
In Fig. 6 the experimental and theoretical data of Fig. 4 are shown again using a logarithmic energy scale in order to focus on a comparison at very low energies. The shape of the experimental spectrum with an additional increase towards low energies $``$ 15 meV is typical for low energy recombination measurements. At E<sub>rel</sub> = 0 eV we obtained a maximum rate coefficient of 1.5$`\times 10^7`$ cm<sup>3</sup> s<sup>-1</sup> exceeding the theoretical rate of 2.9$`\times 10^8`$ cm<sup>3</sup> s<sup>-1</sup> by a factor of 5.2. As already mentioned in the introduction this rate enhancement phenomenon has also been observed at other facilities, however, the Bi<sup>83+</sup>-experiment provides the first quantitative determination of the rate enhancement factor $`ฯต`$ for a bare ion with $`Z18`$. As mentioned above, for the light ions He<sup>2+</sup>, N<sup>7+</sup>, Ne<sup>10+</sup> and Si<sup>14+</sup> a $`Z^{2.8}`$ dependence of $`\mathrm{\Delta }\alpha `$ = $`\alpha _{exp}\alpha _{theo}`$ was found in Stockholm . This scaling cannot be directly confirmed for Bi<sup>83+</sup>. In order to obtain an agreement with the $`Z^{2.8}`$-scaling the measured rate coefficient for Bi<sup>83+</sup> would have to be more than 400 % higher which is beyond the experimental uncertainty. However, one has to be careful with comparing results from different facilities since the experimental conditions vary drastically. Apart from the extremely high electron energy of 162 keV and the extremely high nuclear charge in the present case the influence of the experimental parameters $`kT_{}`$, $`kT_{}`$ and the magnetic guiding field B has to be considered. It is known from previous experiments with F<sup>6+</sup> and C<sup>6+</sup> ions that the excess rate $`\mathrm{\Delta }\alpha `$ scales as $`(kT_{||})^{1/2}`$ and as $`(kT_{})^{1/2}`$. Using the present temperatures in comparison with the Stockholm conditions the excess rate found in the present experiment has to be multiplied by a factor of approximately $`\sqrt{120mev/10meV}\sqrt{0.1meV/0.12meV}`$ 3.2 in order to normalize it to the Stockholm conditions. This removes part of the discrepancy mentioned above. The magnetic field dependence of $`\mathrm{\Delta }\alpha `$ found in similar experiments with other ions would, however, reduce the normalization factor again. To little is known to date about the exact numbers for such a normalization.
As mentioned already in section IV the alignment of the beams has been carefully optimized before starting the recombination experiment. During the measurement we artificially introduced an angle $`\varphi `$ between the beams in order to check the obtained settings. This was implemented by superimposing in the interaction region a defined transverse (horizontal) magnetic field $`B_x`$ in addition to the unchanged longitudinal field $`B_z`$ along the ion beam direction. Fig. 7 shows the maximum recombination rate at $`E_{rel}`$ = 0 eV for different angles $`\varphi `$ from -0.6 mrad to 0.6 mrad in the horizontal plane. At $`\varphi `$ = 0 mrad the maximum recombination rate is obtained. The open circles in Fig. 7 denote the expected rates for the selected angles. They have been determined by taking recombination rates from the ($`\varphi `$ = 0 mrad)-spectrum at the minimum relative energies possible at the corresponding angle (see eq. 16). The squares in Fig. 7 represent the measured rate coefficients at the minimum relative energies $`E_{rel}`$ accessible at the selected angles. All of them are lying above the expected value. Such a behaviour is already known from other experiments at the ESR. Obviously the ion beam reacts on the introduction of the transverse magnetic field component, Lorentz and cooling forces appear to minimize the effect of the change. In any case, the distribution shows a rather symmetric progression with the maximum at $`\varphi `$ = 0 mrad confirming the accurate adjustment of the cooler.
In order to investigate the influence of the electron density on the recombination rate we performed recombination measurements for three different densities $`n_e`$ = 1.6$`\times `$10<sup>6</sup> cm<sup>-3</sup>, $`n_e=3.2\times 10^6`$ cm<sup>-3</sup> and 4.7$`\times `$10<sup>6</sup> cm<sup>-3</sup>. In Fig. 8 the rate coefficient at E<sub>rel</sub>= 0 eV is plotted against the electron density n<sub>e</sub>. The solid line represents the theoretical rate coefficient calculated with kT<sub>||</sub>= 0.1 meV and kT= 120 meV at E<sub>rel</sub>= 0 eV. There is a small difference between the maximum rate coefficient $`\alpha _{max}=1.4\times 10^7`$ cm<sup>3</sup> s<sup>-1</sup> for $`n_e`$ = 1.6$`\times `$10<sup>6</sup> cm<sup>-3</sup> and $`\alpha _{max}=1.5\times 10^7`$ cm<sup>3</sup> s<sup>-1</sup> for the two higher densities but this deviation is within the experimental uncertainty. In addition Fig. 9 shows that the shape of the spectra is equal for all densities indicating identical temperatures in the measurements. Therefore it can be concluded that there is no significant influence of the electron density on the recombination rate. This observation is in accordance with findings at the CRYRING , the TSR and the GSI single pass electron target . The lack of any density dependence rules out TBR (eq. 2) as a possible mechanism leading to enhanced recombination rates at low energies. With a significant contribution of TBR to the observed rates one would expect an increase of the recombination rate with increasing electron density in contrast to all experimental observations. In the context of storage rings TBR has been discussed in some detail by Pajek und Schuch . They found theoretically that TBR effectively populates high Rydberg states of the ion where the electrons are very weakly bound. As mentioned above such ions are reionized in the dipole magnet and therefore do not contribute to the measured recombination rate.
In a next stage of our experiment we also varied the magnetic guiding field of the electron beam between 70 mT and 150 mT in steps of 1 mT. The standard field strength used for the previous measurements was 110 mT. In contrast to the more careful adjustments of the magnetic field at the TSR which were accompanied by measurements of the cooling force and the beam profiles in order to preserve the beam quality no other cooler setting beside the magnetic field was changed at the ESR. This procedure was motivated by an earlier experiment of the ESR cooler group showing an oscillation of the recombination rate for 310 MeV/u U<sup>92+</sup>-ions induced by small changes of the magnetic field. The new results obtained for 295.3 MeV/u Bi<sup>83+</sup> are shown in Fig. 10 where the maximum recombination rate at $`E_{rel}`$ = 0 eV is plotted versus the magnetic field strength $`B_{||}`$. Since the measurement of a complete recombination spectrum is very time-consuming only the recombination rate at cooling has been recorded for each magnetic field setting. For these data points a background subtraction and corrections due to the potential and angle distributions inside the cooler were not possible. Therefore these recombination rates represented by the open circles in Fig. 10 can only display the qualitative dependance on the magnetic field.
A complementary method of determining the recombination rate at cooling can be applied by analyzing the lifetime of the Bi<sup>83+</sup> beam in the ring. Fitting an exponential curve $`I(t)=I_0\mathrm{exp}(t/\tau )+I_{Off}`$ to the decay of the ion current stored in the ring with $`I_0`$ being the ion current at the beginning and with $`I_{Off}`$ representing the offset of the ion current transformer, one can extract the storage lifetime $`\tau `$ of the Bi<sup>83+</sup>-beam in the ring. Assuming that electron-ion recombination in the cooler is the only loss mechanism for stored ions one can calculate the corresponding recombination rate $`\alpha _l=1/(\tau n_{eff})`$. $`n_{eff}`$ is the effective electron density which is $`n_e`$ times the ratio of the interaction length and the ring circumference. Fig. 11 shows a comparison of the rates obtained with the different methods resulting in a good agreement of the data.
A Fourier analysis of the experimental data reveals an โoscillation periodโ of the recombination rate of 7.6 mT. Recent measurements of the ESR cooler group with 300 MeV/u Kr<sup>36+,35+</sup>-ions show that this value can be influenced by a variation of the magnetic fields of the toroids (see Fig. 1) moving the electron beam into and out of the interaction area . Due to the Lorentz force the electrons are moving on helical trajectories through the cooler. It has been observed by the ESR group that the โoscillation periodโ corresponds to a change of the magnetic field that allows the electrons one more turn inside the toroid. In order to find an explanation for these observations one has to take a precise look at the measured recombination spectra. Therefore we performed complete recombination measurements with all corrections for 12 selected field strengths. The results are represented by the full triangles in Fig. 10. The maximum recombination rates at $`E_{rel}`$ = 0 eV obtained in these measurements show the same progression as the open circles.
In Fig. 12 recombination spectra for different magnetic fields between B = 109 mT and 114 mT are shown in the energy range from $`E_{rel}`$ = 0 eV to 125 eV. For $`E_{rel}`$ 1 eV the measured rate coefficients are practically identical. At energies below 1 eV the recombination rates show significant differences depending on the magnetic field strength. However, the rate enhancement is always there which one can see in Fig. 13 where the recombination rate for the oscillation minimum at B = 114 mT is compared with RR theory. In order to describe the experimental data a transverse electron temperature of $`kT_{}`$ = 450 meV was applied which is nearly a factor of 3 higher than the one obtained by a fit of RR theory to the experimental data at the standard magnetic field strength of B = 110 mT. From the measured rate at $`E_{rel}`$ = 0 eV of $`\alpha _{exp}`$ = 6.6$`\times 10^8`$ cm<sup>3</sup> s<sup>-1</sup> and the theoretical value at $`E_{rel}`$ = 0 eV of $`\alpha _{theo}`$ = 1.5$`\times 10^8`$ cm<sup>3</sup> s<sup>-1</sup> one calculates a rate enhancement factor $`ฯต=\alpha _{exp}/\alpha _{theo}`$ = 4.4.
Analysing all the complete recombination spectra measured for different magnetic fields a connection between the maximum rate coefficient at $`E_{rel}`$ = 0 eV and the adapted transverse electron temperature $`kT_{}`$ becomes obvious which is documented in Fig. 14. In the left panel representing the data for magnetic fields between 76 mT and 80 mT the measured maximum rate coefficient at $`E_{rel}`$ = 0 eV (full triangles) increases with increasing magnetic field whereas the transverse electron temperature (open circles) adapted to the corresponding recombination spectra decreases in this region. In the right spectrum monitoring the same data for magnetic fields between 109 mT and 114 mT one can see a decrease of the maximum rate coefficient at $`E_{rel}`$ = 0 eV accompanied by an increase of the transverse electron temperature $`kT_{}`$. Therefore there seems to be a relationship between the rate coefficient and the transverse electron temperature (or: the electron beam quality) which would easily explain the periodic reductions of the recombination rate at $`E_{rel}`$ = 0 eV. Nevertheless it should be pointed out that this interpretation does not explain the observed general enhancement of the rate coefficient at $`E_{rel}`$ = 0 eV. A further theoretical analysis should especially focus on the possible influence of the electron energy since in experiments with lithium-like 97.2 MeV/u Bi<sup>80+</sup> ions (corresponding to 53.31 kV cooling voltage instead of 162 kV for 295.3 MeV/u) during the same beam-time the oscillations did not appear. This is in agreement with the results obtained at the ESR and other storage rings in experiments at low ion energies. There, oscillations of the recombination rate at cooling have not been observed.
## VI Conclusions
The recombination of bare Bi<sup>83+</sup> ions with free electrons has been studied at the GSI Experimental Storage Ring (ESR) in Darmstadt. Within the experimental uncertainty we found very good agreement between the measured rate coefficient and the theory for radiative recombination (RR) for energies from E<sub>rel</sub> = 15 meV to 125 eV. At very low center-of-mass energies between ions and electrons, however, the measured rate exceeds the theoretical predictions by a factor of 5.2. This first rate enhancement result for a bare ion with $`Z`$ very much greater than 18 does not follow a $`Z^{2.8}`$ dependence of the enhancement found in an experiment with the light ions He<sup>2+</sup>, N<sup>7+</sup>, Ne<sup>10+</sup> and Si<sup>14+</sup>. Therefore additional measurements in the mid-Z range are necessary in order to obtain more information about the influence of $`Z`$ on the recombination enhancement phenomenon. The increase of the electron density in the interaction area from $`n_e`$ = 1.6$`\times `$10<sup>6</sup> cm<sup>-3</sup> to 4.7$`\times `$10<sup>6</sup> cm<sup>-3</sup> appears to have no significant effect on the recombination rate. A variation of the magnetic field from 70 mT to 150 mT revealed a strong dependence of the recombination rate at low energies on this parameter. The observed oscillations of the maximum recombination rate at $`E_{rel}`$ = 0 eV confirmed previous observations with bare U<sup>92+</sup> ions. Comparing the recombination spectra with RR theory one finds strong variations of the transverse electron temperature connected to the oscillations of the recombination rate. In future experimental and theoretical studies this relationship has to be investigated in more detail. Finally, we want to emphasize, that the observed recombination rate enhancement significantly reduces the lifetime of ion beams in storage rings during the electron cooling procedure. At the present ion energies, recombination in the cooler by far dominates over all factors influencing the beam lifetime. The recombination rate enhancement hence reduces the beam lifetime by approximately a factor 5 as compared to the assumption of pure RR.
## VII Acknowlegments
The Giessen group gratefully acknowledges support for this work through contract GI MรL S with the Gesellschaft fรผr Schwerionenforschung (GSI), Darmstadt, and by a research grant (number 06 GI 848) from the Bundesministerium fรผr Bildung und Forschung (BMBF), Bonn. |
warning/0003/nucl-th0003050.html | ar5iv | text | # Pion-nucleon interaction in a covariant hadron-exchange model
## I Introduction
The $`\pi N`$ interaction received much attention in the past both theoretically and experimentally in view of its fundamental nature (early literature can be found in Refs. ). Current theoretical interest is triggered by the experimental programs being carried out at NIKHEF, MAMI, TJNAF and other intermediate-energy facilities with the purpose to understand the structure of hadrons and their interaction in the confinement region of QCD. To extract most physics from the new high-precision measurements a reliable and accurate knowledge of $`\pi N`$ and $`NN`$ interaction is required. Highly successful attempts have been made in describing these interactions in terms of hadronic degrees of freedom over a wide energy region. In particular, the relativistic one-boson-exchange models were successfully applied to the description of $`NN`$ interaction, and especially during the past decade this theoretical framework has been extended to the $`\pi N`$ system . Such an extension gives one a capability to study a broad class of reactions, including pion scattering and production on light nuclei in a self-consistent framework. In this paper we report on a relativistic covariant $`\pi N`$ model using the quasipotential approach and based on an effective interaction characterized by a hadron-exchange potential (some of our results have already been briefly reported ).
Although the underlying dynamics of the $`\pi N`$ interaction is nowadays believed to be governed by QCD, it is practically impossible to resolve it fully in terms of quarks and gluons because of the confinement problem. Much of the present understanding of the $`\pi N`$ physics at low and intermediate energies remains to be based on dispersion relations and effective chiral Lagrangians in term of the hadron degrees of freedom.
The chiral pion-nucleon Lagrangians are usually extended in two ways: first, by including higher-mass states, such as $`\rho `$-meson, $`\mathrm{\Delta }`$-isobar, etc.; secondly, by including the higher-derivative terms. Both ways are necessary to extend such a phenomenological description to higher energies: The contributions due to higher-mass states have a clear physical significance, while the higher-derivative terms are needed to examine the effect of unknown short-range physics. The higher-derivative terms play, for instance, a crucial role in the renormalization program of chiral perturbation theory (ChPT). In the hadron-exchange models, a similar role is played by the โstrong form factorsโ which are included in the effective Lagrangian to model the short distance behavior of the potential. Both ChPT and hadron-exchange models thus begin from a similar โextendedโ chiral Lagrangian but the approach to calculating the $`\pi N`$ scattering amplitude is somewhat different. In ChPT one usually performs a perturbative field-theoretic calculation maintaining crossing-symmetry and exact agreement with the soft-pion theorems. (For the development of ChPT in application to the $`\pi N`$ scattering see Ref. .) In the hadron-exchange approach one uses the effective Lagrangian to construct the potential which then is resumed via a scattering equation. In this way crossing symmetry is given up in favor of exact unitarity in a given channel space, and possibility of studying nonperturbative phenomena such as dynamical resonances.
In defining a hadron-exchange model one usually specifies three ingredients: effective Lagrangian, potential, and the scattering equation. These ingredients are interrelated in quantum field theory, where one must solve the Bethe-Salpeter (BS) equation and has a well-determined procedure for computing its kernel from a given Lagrangian. The BS kernel consists of all the irreducible graphs and hence can be computed only perturbatively. In this work we shall take the potential to be given by the tree-level graphs, although we do not claim that the used Lagrangian justify any perturbative expansion. On the other hand, the resulting approximation transparently relates to the usual quantum-mechanical picture where the scattering problem is given by a Lippmann-Schwinger type of equation for one-particle exchange potentials. Therefore, one might prefer to view this approach as relativistic quantum-mechanical one rather than some โnonsystematicโ truncation of QFT.
The four-dimensional BS equation for the $`\pi N`$ system (with a one-particle-exchange potential) has been solved by Nieland and Tjon , and recently by Lahiff and Afnan in a more realistic setup. Models exploit instead various quasipotential (QP) equations, which can be obtained by a three-dimensional reduction of the BS equation. The use of QP equations provides a technical simplification of the problem, without destroying the Lorentz invariance of the theory. It should be remarked, however, that some of the QP equations can violate charge conjugation symmetry, and because this symmetry is crucial for renormalizing the positive- and negative-energy baryon poles in an equivalent way, the equations which violate it are less preferable. We will employ the equal-time (ET) equation which preserves the full Lorentz covariance, including charge conjugation. This equation will be specified in the next section.
Apart from the technical simplifications, the QP approach can sometimes be motivated by physical arguments. For instance, while the four-dimensional BS equation for $`t`$-channel (meson-exchange) potential, i.e., the ladder BS equation, has a wrong one-body limit, a number of QP equations with the proper limit can be devised . In the $`\pi N`$ situation the potential may, in addition to the $`t`$-channel meson exchanges, contain the $`u`$-channel baryon exchanges, which spoil the standard one-body limit arguments . The ordinary ET equation is shown to provide an optimal choice in the case when both $`t`$\- and $`u`$-channel exchanges are present in the force corresponding to the $`\pi N`$ situation.
As for the effective Lagrangian, we use the pseudovector $`\pi NN`$ coupling, and include $`\sigma `$, $`\rho `$ mesons, and all the relevant (for the considered energy region) nucleon resonances as explicit degrees of freedom. The precise form of the Lagrangian and the potential is discussed in Sec. IV. An interesting aspect, which comes in with the nonperturbative modelling, is that for a sufficiently attractive potential the nucleon resonances can be generated dynamically, as quasi-bound states of the $`\pi N`$ system. In models and quantum hadrodynamics the $`\mathrm{\Delta }`$(1232) is described in this way. Lahiff and Afnan include the $`\mathrm{\Delta }`$ explicitly, but suggest that the Roper resonance can be of dynamical origin. Gross and Surya include the $`\mathrm{\Delta }`$ and the Roper poles, but treat the $`S_{11}`$ resonance dynamically. In this paper we consider $`P_{33}\mathrm{\Delta }`$(1232), $`P_{11}N^{}`$(1450), $`D_{13}N^{}`$(1520), and $`S_{11}N^{}`$(1535) resonances. Within our model, these resonances are all of nondynamical origin, i.e., are included explicitly via an effective Lagrangian description. Of course, the dynamical effects will anyhow contribute to the generation of the resonances seen in the phase-shifts. Thus, an admixture of both โelementaryโ and โcompositeโ component constitutes the full result. Since the โelementaryโ fields corresponding to the resonances are included with real masses, the dynamical contributions are fully responsible for generating the width. Our model maintains the elastic $`\pi N`$ unitarity and therefore only the one-pion decay width of the resonances is generated.
As we have to deal with the spin-3/2 fields of resonances, such as that of the $`\mathrm{\Delta }`$-isobar, we shall address here some of the problems of consistent formulation for relativistic higher-spin fields. Consistent formulations for the free spin-3/2 field have of course been known for a long time. The Rarita-Schwinger formalism based on the vector-spinor field representation became the most popular one. The form of the free spin-3/2 action is uniquely (up to trivial field redefinitions) constrained by requirements of Poincarรฉ invariance and consistent degrees-of-freedom counting. The latter requirement essentially means that the action must have enough symmetries to kill off the unphysical lower-spin components, and maintain only the physical $`2s+1`$ degrees-of-freedom of the theory. An arbitrarily constructed interaction may violate this consistency and activate the unphysical degrees of freedom. This necessarily leads to a number of pathologies, such as acausal propagations, inadmissible quantization, violation of Poincarรฉ invariance.
The conventional $`\pi N\mathrm{\Delta }`$ coupling is an example of such pathological interactions. The contribution of the unphysical spin-1/2 components appear in the scattering amplitudes via the dependence on the so-called โoff-shell parameterโ and โspin-1/2 backgrounds.โ On the other hand, a class of consistent $`\pi N\mathrm{\Delta }`$ and $`\gamma N\mathrm{\Delta }`$ couplings has recently been established . Those are essentially all possible couplings that maintain the gauge symmetry of the free massless Rarita-Schwinger action. In the present model we shall use the leading โgauge-invariantโ $`\pi N\mathrm{\Delta }`$ coupling, which has the same non-relativistic limit as the conventional one. In Sec. V we study the differences between the conventional and the gauge-invariant $`\pi N\mathrm{\Delta }`$ coupling at the tree-level $`\mathrm{\Delta }`$-exchange contributions. The largest differences are seen first of all in the spin-1/2 partial waves, where the conventional coupling gives the background contributions verse no contribution from the consistent couplings.
The parameters of the effective Lagrangian, including the form factor masses, form the set of model parameters. Some of them, such as the $`\pi NN`$ coupling constant, the nucleon, and the meson masses, are very well-determined elsewhere and therefore are kept fixed during the fits. The others are fitted to give the best agreement with the $`\pi N`$ partial-wave analyses . The complete model provides an accurate description of the $`S`$\- and $`P`$-wave scattering lengths ($`\chi ^2/\mathrm{datapoint}1.4`$), as well as the energy behavior of the $`S`$-, $`P`$-, and some of the $`D`$-wave phase-shifts up to 600 MeV pion lab kinetic energy.
In the next section we describe the covariant quasipotential equation for the off-shell $`\pi N`$ amplitudes, and its partial-wave decomposition. This will specify the equation solved in the model. In Sec. III we briefly discuss the effective Lagrangian used to read off the tree-level potentialโthe driving force of the equation. The renormalization procedure to treat the s-channel poles is described in Sec. IV. In Sec. V we analyze the effects of the different exchange contributions in the low-energy $`\pi N`$ data, and then present the results of the complete model emphasizing the effect of the rescattering contributions. Some discussion and concluding remarks are given in Sec. VI. Finally, various appendices contain some technical details of the analysis. In Appendix A we summarize the conventions. Appendix B shows some details of the partial-wave and isospin decomposition of the off-shell $`\pi N`$ amplitudes. Appendix C provides explicit expressions for various hadron-exchange contributions to the off-shell $`\pi N`$ potential. Amplitudes for higher-spin baryon exchange are discussed in Appendix D.
## II Quasipotential approach
The fully off-shell relativistic $`\pi N`$ scattering amplitude in the space of the nucleon helicity spinors is described by sixteen scalar amplitudes: one for each combination of the helicity and $`\rho `$-spin of the initial and final nucleon. Parity conservation reduces the number of independent scalar amplitudes to eight. As a suitable covariant representation which expresses the off-shell amplitude in terms of eight invariants we choose the following:
$$T_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},k^{};p,k)=\overline{u}_\lambda ^{}^\rho ^{}(\stackrel{}{p}^{})(1,p/^{})[\left(\begin{array}{cc}A_{11}& A_{12}\\ A_{21}& A_{22}\end{array}\right)+P/\left(\begin{array}{cc}B_{11}& B_{12}\\ B_{21}& B_{22}\end{array}\right)]\left(\begin{array}{c}1\\ p/\end{array}\right)u_\lambda ^\rho (\stackrel{}{p}),$$
(1)
where $`A_{ij}`$ and $`B_{ij}`$ are the eight scalar functions of invariants formed by the momenta, i.e., $`p^2`$, $`pp^{}`$, etc. The four-momenta of the initial (final) nucleon and pion are given by $`p`$ and $`k`$ ($`p^{}`$ and $`k^{}`$), respectively, while $`P`$ is the conserved total four-momentum of the system:
$$P=p+k=p^{}+k^{}.$$
Due to the momentum conservation only three of the external momenta are independent, below we usually work with $`p`$, $`p^{}`$ and $`P`$. Furthermore, $`u_\lambda ^\rho `$ are the nucleon helicity spinors, where $`\lambda =\pm \frac{1}{2}`$ and $`\rho =\pm 1`$ ($`\lambda ^{}`$ and $`\rho ^{}`$) are the initial (final) helicity and $`\rho `$-spin of the nucleon, respectively.
For the on-shell situation ($`p^2=p_{}^{}{}_{}{}^{2}=m_N^2`$, $`k^2=k_{}^{}{}_{}{}^{2}=m_\pi ^2`$) the amplitude reduces to the standard form :
$$T_{\lambda ^{}\lambda }(p^{},p;P)=\overline{u}_\lambda ^{}^{(+)}(\stackrel{}{p}^{})\left[\widehat{A}(s,t)+\frac{1}{2}\gamma (k^{}+k)\widehat{B}(s,t)\right]u_\lambda ^{(+)}(\stackrel{}{p}),$$
(2)
\[$`s=P^2,t=(pp^{})^2,u=(pk^{})^2`$ are the Mandelstam invariants\]. We find from Eqs. (1) and (2)
$`\widehat{B}`$ $`=`$ $`B_{11}+m_N(B_{12}+B_{21})+m_N^2B_{22},`$ (3)
$`\widehat{A}`$ $`=`$ $`m_N\widehat{B}+A_{11}+m_N(A_{12}+A_{21})+m_N^2A_{22}.`$ (4)
Our starting point for the $`\pi N`$ amplitude is the Bethe-Salpeter (BS) equation, schematically shown in Fig. 1,
$$T(p^{},p)=V(p^{},p)+i\frac{\mathrm{d}^4q}{4\pi ^3}V(p^{},q)G(q)T(q,p),$$
(5)
where $`V`$ is the potential, $`G`$ is the $`\pi N`$ propagator; the dependence on the total momentum $`P`$ is omitted. If $`q`$ is the relative four-momentum of the intermediate $`\pi N`$ state, the $`\pi N`$ propagator takes the following form:
$$G(q)=\frac{1}{(\beta Pq)^2m_\pi ^2+i\epsilon }\frac{(\alpha P+q)\gamma +m_N}{(\alpha P+q)^2m_N^2+i\epsilon },$$
(6)
where
$`\alpha \alpha (s)=pP/s=(s+m_N^2m_\pi ^2)/2s,`$ (7)
$`\beta \beta (s)=kP/s=(sm_N^2+m_\pi ^2)/2s.`$ (8)
In approximating the BS equation one often simplifies the singularity structure of the kernel $`VG`$, such that the temporal integration can easily be done. This procedure is called three-dimensional (3D) reduction while the resulting equation is a quasipotential (QP) equation. For instance, in the reduction to the spectator equation all the poles of $`V`$ and the negative-energy pole of $`G`$ in the $`q_0`$ plane are neglected.
As we have recently emphasized , the danger of doing a 3D reduction via approximating the pole structure is that the charge conjugation symmetry can be destroyed. In particular, in our naive interpretation of the spectator equation this symmetry is violated, essentially because of an asymmetric treatment of the positive- and negative-energy states. Gross has recently presented an interpretation of the spectator prescription which is consistent with the charge conjugation symmetry .
The equal-time (ET) reductions (see, e.g. ), such as Salpeterโs instantaneous approximation , preserve charge conjugation symmetry. In these reductions one effectively removes the $`q_0`$ poles from the potential while treating exactly the poles of the two-particle propagator $`G`$. To remove the potential poles one fixes the relative-energy variable $`q_0`$ in some way. Most frequently the constraint $`q_0=0`$, or its covariantized form: $`Pq=0`$, is used.
We will be using the ET type of approach. To implement the constraint $`Pq=0`$, we may impose the condition that the interaction is insensitive to the off-shellness along the direction defined by an unit four-vector $`n_\mu `$. For the two-body case this means that $`V`$ and $`T`$ entering the scattering equation depend on the projections of the relative four-vectors onto a 3D hyperplane orthogonal to $`n_\mu `$. Defining the projection operator,
$$O_{\mu \nu }=\mathrm{g}_{\mu \nu }n_\mu n_\nu ,$$
(9)
we write the corresponding equation as follows:
$$T(\stackrel{~}{l}^{},\stackrel{~}{l})=V(\stackrel{~}{l}^{},\stackrel{~}{l})+i\frac{\mathrm{d}^4q}{4\pi ^3}V(\stackrel{~}{l}^{},\stackrel{~}{q})G(q)T(\stackrel{~}{q},\stackrel{~}{l}),$$
(10)
where $`l,l^{},q`$ are the relative momenta of the initial, final, and intermediate $`\pi N`$ state, respectively; $`\stackrel{~}{l}_\mu =O_{\mu \nu }l^\nu `$, and similarly for $`\stackrel{~}{l}^{},\stackrel{~}{q}`$.
Equation (10) is manifestly covariant. On the other hand, it can easily be reduced to the 3D form. For instance, we can choose the frame where $`n=(1,0,0,0)`$, and therefore $`V`$ and $`T`$ are independent of the 0-th component of relative momenta (since any scalar product will depend only on the spatial components, e.g., $`\stackrel{~}{q}\gamma =\stackrel{}{q}\stackrel{}{\gamma }`$). The integration over $`q_0`$ in Eq. (10) can now be readily done leading to the 3D equation. To prevent the dependence of the $`S`$-matrix on $`n`$, one may choose $`n`$ along some physical four-momentum, for instance along the total momentum: $`n_\mu =P_\mu /\sqrt{P^2}.`$ Then the reduction is possible in the center-of-mass system (CMS) where $`P=(P_0,0,0,0)`$.
In the ET approach, the two-particle propagator is sometimes modified to include approximately the crossed graphs , thus providing the correct one-body limit of the equation in the case of $`t`$-channel type of potential. We however do not apply such modifications here, because they actually worsen the predictions for the $`\pi N`$ case where the $`u`$-channel exchanges are present .
Because of rotational invariance and parity conservation it is convenient to partial-wave decompose Eqs. (5) and (10) (see Appendix B). Let
$$T_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p)=_J(J+\frac{1}{2})D_{\lambda ^{}\lambda }^J(\mathrm{\Omega }_{p^{}p})T_{\lambda ^{}\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0),$$
(11)
where $`\mathrm{\Omega }_{p^{}p}`$ is the solid angle between $`\stackrel{}{p}^{}`$ and $`\stackrel{}{p}`$. Furthemore, in the CMS, using Eq. (1) and the Dirac equation
$$(p/\rho m_N)u_\lambda ^\rho (\stackrel{}{p})=(p_0\rho E_p)\gamma _0u_\lambda ^\rho (\stackrel{}{p}),$$
(12)
the off-shell amplitudes can be written as
$$T_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p;P_0)=\overline{u}_\lambda ^{}^\rho ^{}(\stackrel{}{p}^{})\left[\gamma _+T_+^{\rho ^{}\rho }(p^{},p;P)+\gamma _{}T_{}^{\rho ^{}\rho }(p^{},p;P_0)\right]u_\lambda ^\rho (\stackrel{}{p}),$$
(13)
where $`\gamma _\pm =\frac{1}{2}(I\pm \gamma _0)`$, and $`T_\pm ^{\rho ^{}\rho }`$ are eight scalar amplitudes with definite parity $`\pm `$.
In doing so we in particular find for the case of the ET equations, that the parity-conserving amplitudes $`T_r^J`$ satisfy
$$T_r^{J\rho ^{}\rho }=V_r^{J\rho ^{}\rho }+\frac{1}{\pi }\underset{0}{\overset{\mathrm{}}{}}dqq^2\underset{\rho ^{\prime \prime }}{}G_{ET}^{(\rho ^{\prime \prime })}V_r^{J\rho ^{}\rho ^{\prime \prime }}T_r^{J\rho ^{\prime \prime }\rho },$$
(14)
where
$`G_{ET}^{(\rho )}(|\stackrel{}{q}|;P_0)`$ $`=`$ $`2i{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{d}q_0}{2\pi }}G^{(\rho )}(q,q_0;P_0)`$ (15)
$`=`$ $`\rho \left\{\omega _q(\rho P_0+E_q+\omega _qiฯต)\right\}^1.`$ (16)
In this work we will be focusing on solving this equation for a one-particle-exchange $`\pi N`$ potential, see Fig. 2, where the potential is regulated by form factors. In the next section we describe the $`\pi N`$ interaction used in this study.
## III Effective Lagrangian and the $`\pi N`$ potential
In the following we specify the interaction Lagrangian of $`\pi `$, $`\rho `$, $`\sigma `$, $`N`$ and isobar fields, used to construct the hadron-exchange force depicted in Fig. 2 and written out in Appendix C. The field representation and the corresponding free Lagrangian is chosen according to the spin and isospin of the particle. Thus, the pion is described by scalar isovector multiplet $`\pi ^a=(\pi ^+,\pi ^0,\pi ^{})`$, the sigma-meson is a scalar isoscalar field $`\sigma `$, the $`\mathrm{\Delta }`$(1232) is represented by a vector-spinor isoquadruplet $`\mathrm{\Delta }_\mu =(\mathrm{\Delta }_\mu ^{++},\mathrm{\Delta }_\mu ^+,\mathrm{\Delta }_\mu ^0,\mathrm{\Delta }_\mu ^{})`$, and so on.
### A Nucleon and meson exchanges
The $`\pi NN`$ interaction Lagrangian is taken in accordance with the chirally-symmetric $`\sigma `$ models. In Weinbergโs nonlinear realization the $`\pi N`$ scattering amplitude to the leading order is given by the nucleon Born term with the pseudovector coupling plus the Weinberg-Tomozawa contact term . The pseudovector coupling reads
$$_{\pi NN}^{(\mathrm{PV})}=\frac{f_{\pi NN}}{m_\pi }\overline{N}\gamma _\mu \gamma _5\tau _aN_\mu \pi ^a,$$
(17)
where $`f_{\pi NN}`$ is the pseudovector $`\pi NN`$ coupling constant (the pseudoscalar coupling constant: $`g_{\pi NN}=2f_{\pi NN}(m_N/m_\pi )`$ will also be used below). The Weinberg-Tomozawa contact term can be represented as a $`\rho `$-meson exchange with the following interaction:
$`_{\rho NN}`$ $`=`$ $`g_{\rho NN}\overline{N}{\displaystyle \frac{\tau _a}{2}}\left(\gamma _\mu \rho _a^\mu +{\displaystyle \frac{i\kappa _\rho }{2m_N}}\sigma _{\mu \nu }^\mu \rho _a^\nu \right)N,`$ (18)
$`_{\rho \pi \pi }`$ $`=`$ $`g_{\rho \pi \pi }\epsilon _{abc}\rho _a^\mu \pi _b_\mu \pi _c,`$ (19)
provided the $`\rho `$ coupling, $`g_\rho ^2=g_{\rho NN}g_{\rho \pi \pi }`$, is fixed by the KSRF relation : $`g_\rho =m_\rho /(\sqrt{2}f_\pi )`$, where $`f_\pi 93`$ MeV. There is also the second form of the KSRF relation: $`g_\rho =\sqrt{2}m_\rho f_{\pi NN}/(m_\pi g_\mathrm{A})`$, $`g_\mathrm{A}1.26`$, obtained from the first one by using the Goldberger-Triemann relation for $`f_\pi `$. It should be remarked that the Weinberg-Tomozawa contact term is equivalent to the $`\rho `$-exchange only at threshold and provided $`g_\rho `$ is fixed by KSRF relation while $`\kappa _\rho =0`$. The energy dependence is different, but not significantly in the considered energy region.
Since we use the pseudovector $`\pi NN`$ coupling, the $`\sigma `$-meson is in principle not needed from the standpoint of chiral symmetry. Nevertheless, a $`\sigma `$ exchange can be used to model the isoscalar contribution of the correlated two-pion exchange. In order to keep the agreement with the soft-pion theorem, a derivative coupling to the pion is used, i.e.,
$`_{\sigma NN}`$ $`=`$ $`g_{\sigma NN}\sigma \overline{N}N,`$ (20)
$`_{\sigma \pi \pi }`$ $`=`$ $`{\displaystyle \frac{g_{\sigma \pi \pi }}{2m_\pi }}\sigma _\mu \pi _i^\mu \pi _i,`$ (21)
where the sign of the $`\sigma \pi \pi `$-coupling is chosen in accordance with the correlated two-pion exchange analysis, and is different from the one used in . This interaction leads to the following on-shell potential:
$$V_\sigma =\frac{g_{\sigma NN}g_{\sigma \pi \pi }}{8\pi m_\pi }\frac{t2m_\pi ^2}{m_\sigma ^2t}.$$
(22)
To control the effect of the $`\sigma `$-exchange on the scattering length we introduce a free parameter $`c_\sigma `$ in the following way:
$$V_\sigma =\frac{g_{\sigma NN}g_{\sigma \pi \pi }}{8\pi m_\pi }\frac{t2(1c_\sigma )m_\pi ^2}{m_\sigma ^2t}.$$
(23)
For $`c_\sigma =1`$ the $`\sigma `$ contribution to the $`S`$-wave scattering lengths vanishes. Note that this modification amounts to adding the following term to the Lagrangian:
$$_{\sigma \pi \pi }^{}=\frac{g_{\sigma \pi \pi }}{2}c_\sigma m_\pi \sigma \pi _i^2.$$
(24)
### B $`\mathrm{\Delta }`$-isobar exchange and higher resonances
The coupling of the spin-3/2 $`\mathrm{\Delta }`$ field to the pion and the nucleon is conventionally described by the following Lagrangian (see, e.g., Ref. ):
$$_{\pi N\mathrm{\Delta }}=\frac{f_{\pi N\mathrm{\Delta }}}{m_\pi }\overline{\mathrm{\Delta }}^\mu \left[\mathrm{g}_{\mu \nu }(z+\frac{1}{2})\gamma _\mu \gamma _\nu \right]T_aN_\nu \pi ^a+\text{H.c.},$$
(25)
where $`z`$ is the off-shell parameter, $`T_a`$ is the isospin-$`\frac{1}{2}\frac{3}{2}`$ transition operator.
As remarked in the Introduction, this coupling involves the spin-1/2 sector of the Rarita-Schwinger field, and give rise to an unphysical spin-1/2 background. The latter effect can in principle be removed by inserting the spin-3/2 projection operator โby handโ in either the vertex or the propagator. For instance, Gross and Surya have chosen this option. However, because of the nonlocal nature of the projection operators their use is problematic: unphysical singularities occur at $`s=0`$ and $`u=0`$ for the $`s`$\- and $`u`$-channel contribution, respectively. In Ref. this problem is actually not met because the $`s=0`$ point is well below the threshold, while the $`u`$-graph vanishes in the approximation of that model. If the $`u`$-channel $`\mathrm{\Delta }`$ exchange is present, one usually prefers to keep the background and fit the off-shell parameter .
We shall also study the following $`\pi N\mathrm{\Delta }`$ coupling :
$$_{\pi \mathrm{N}\mathrm{\Delta }}^{(\mathrm{GI})}=\frac{f_{\pi N\mathrm{\Delta }}}{m_\pi m_\mathrm{\Delta }}\epsilon ^{\mu \nu \alpha \beta }(_\mu \overline{\mathrm{\Delta }}_\nu )\gamma _5\gamma _\alpha T_aN_\beta \pi ^a+\text{H.c.},$$
(26)
referred to as gauge-invariant (GI) $`\pi N\mathrm{\Delta }`$ coupling. Being invariant under the Rarita-Schwinger gauge transformation:
$$\mathrm{\Delta }_\mu (x)\mathrm{\Delta }_\mu (x)+_\mu ฯต(x),$$
where $`ฯต(x)`$ is a spinor field, this coupling does not involve the spin-1/2 components of the $`\mathrm{\Delta }`$ field. As a consequence, the spin-1/2 backgrounds are totally absent from the corresponding $`\mathrm{\Delta }`$-exchange amplitudes.
We include also the $`s`$\- and $`u`$-channel graphs of $`P_{11}`$ (Roper), $`S_{11}`$, and $`D_{13}`$ resonances. At low energies the contribution of these resonances is marginal,<sup>*</sup><sup>*</sup>*This is generally not true for a spin-3/2 resonance if the conventional coupling is used, since the spin-1/2 background can be large even far away from the mass position. but they are important for the proper description at higher energies. The first two particles are treated same as the nucleon but with different masses, coupling parameters, and, in the case of $`S_{11}`$, parity. The $`D_{13}`$ is treated in the same way as the $`\mathrm{\Delta }`$ (the same propagator and interaction vertex), but with different isospin, parity and mass. Exchanges of even higher spin resonances can in principle be easily included in our model via the amplitude obtained in Appendix D.
### C Cutoff form factors
The high-energy behavior of the hadron-exchange field-theoretic potential is usually regulated using the off-shell form factors introduced in the vertices. We have introduced them for each of the particle in the vertex. For the pion we use the monopole form factor:
$$f_\pi (k^2)=\frac{\Lambda _\pi ^2m_\pi ^2}{\Lambda _\pi ^2k^2}.$$
(27)
For the $`\sigma `$\- and $`\rho `$-meson we use the one-boson-exchange form factor:
$$f_{\sigma ,\rho }(t)=\frac{\Lambda _{\sigma ,\rho }^2}{\Lambda _{\sigma ,\rho }^2t}.$$
(28)
For the baryons we use the form factor of Ref. :
$$f_B(p^2)=\left(\frac{n\Lambda _B^4}{n\Lambda _B^4+(p^2m_B^2)^2}\right)^n$$
(29)
with $`n=2`$.
In addition, for each pion we introduce the following cutoff:
$$f_{Regge}(q,s)=\frac{\Lambda _\pi ^4}{\Lambda _\pi ^4+s\stackrel{}{q}^2}.$$
(30)
This function is motivated by considering the effect of the higher-mass states on the high-energy behavior of the $`\pi N`$ propagator $`G`$. If we were to include not only the pion and the nucleon but all the states lying on the infinitely rising Regge trajectories, at very high energies their effect factorizes in the form of function like Eq. (30) . The energy dependence and the absence of singularities distinguish this cutoff from the usual ones, such as the monopole form. We take the same value $`\Lambda _\pi `$ for the $`f_\pi `$ and $`f_{Regge}`$ cutoff masses. These two form factors do not affect the on-shell potential, because no pion exchanges appear in the Born graphs.
It is important to realize that the final results depend on the off-shell form factors, even after the renormalization is applied. The physical meaning of such form factors is usually given in analogy with that of the electromagnetic form factors. They thus reflect the extension of the hadrons, and in principle should be calculated from the underlying theory.
Our fit to the $`\pi N`$-scattering phases will determine the values of the cutoff masses. They are given in Table V together with the rest of the model parameters. Using these values, in Fig. 3 we have plotted the form factors which affect the loop contributions. Their dependence on the loop momentum is shown, while the 0-th component is fixed by the equal-time constraint and the energy is fixed at threshold.
The actual cutoff of the model is given by the solid line in the figure. As one can see, it is rather soft: it starts off as a monopole with the mass about 0.8 GeV, and is even softer above $`q^2=0.5`$ GeV<sup>2</sup>. At higher energy it becomes softer as well, because $`f_{Regge}`$ is energy-dependent. However, the latter effect is small as can be seen from Fig. 4, where the energy dependence of $`f_{Regge}`$ is shown over the region of our $`\pi N`$ fit. The energy dependence of $`f_N(s)`$ and $`f_\mathrm{\Delta }(s)`$ is shown there as well.
## IV Renormalization
Since there are s-channel singularities in the considered potential, we have to carry out a renormalization procedure. We adopt the scheme in which the Lagrangian is expressed in terms of the physical parameters and no โbareโ parameters appear. Then, in principle, the counter-terms should be subtracted and fixed by the renormalization conditions. To perform such a renormalization procedure it is convenient to work with the one-particle-irreducible Green functions. One can separate the potential into two terms $`V=V_s+V_u`$, where
$$V_s(p^{},p)=\underset{B}{}\Gamma _B(p^{})S_B(P)\Gamma _B(p)$$
(31)
represents the $`s`$-channel baryon exchanges (pole terms), while $`V_u`$ contains the rest of the graphs (nonpole terms). Since $`V_sG`$ is a separable kernel, we can explicitly resum these contributions, and find that the resulting amplitude can equivalently be written as
$$T(p^{},p)=\underset{BB^{}}{}\Gamma _B^{}^{}(p^{})S_{BB^{}}^{}(P)\Gamma _B^{}(p)+T_u(p^{},p),$$
(32)
where
$`\Gamma _B^{}(p)`$ $`=`$ $`\Gamma _B(p)+i{\displaystyle \frac{\mathrm{d}^4q}{4\pi ^3}\Gamma _B(q)G(q)T_u(q,p)},`$ (33)
$`(S_{BB^{}}^{})^1`$ $`=`$ $`(S_B^{})^1\delta _{BB^{}}\Sigma _{BB^{}},`$ (34)
$`\Sigma _{BB^{}}`$ $`=`$ $`i{\displaystyle \frac{\mathrm{d}^4q}{4\pi ^3}\Gamma _B(k)G(k)\Gamma _B^{}^{}(q)},`$ (35)
and $`T_u`$ satisfies the following integral equation,
$$T_u(p^{},p)=V_u(p^{},p)+i\frac{\mathrm{d}^4q}{4\pi ^3}V_u(p^{},q)G(q)T_u(q,p).$$
(36)
The full amplitude is thus written in terms of the irreducible Green functions: $`S^{},\Gamma ^{}`$ and $`T_u`$. The diagrammatic form of this representation is given in Fig. 5.
### A Baryon mixing
Note that the dressed baryon propagator, Eq. (34), is in general nondiagonal. In other words, the baryons can mix. Of course this mixing happens only among the baryons with the same โgoodโ quantum numbers, such as spin and isospin. Parity is also conserved, nevertheless the mixing of baryons with the same spin, isospin and opposite parity may occur due to the negative-energy state propagation.
To perform the renormalization we need first to diagonalize the propagator. Since it is a complex symmetric matrix, we diagonalize it using a complex orthogonal transformation $`๐ช`$ ($`๐ช๐ช^T=๐ช^T๐ช=`$1). The full solution can obviously be written in the diagonal form as follows:
$$T=\underset{B}{}\left(\Gamma ^{}๐ช\right)_B\left(๐ช^TS^{}๐ช\right)_B\left(๐ช^T\Gamma ^{}\right)_B+T_u,$$
(37)
In our model we include only two baryons with the same spin and isospin (nucleon and $`N^{}`$). For this case the propagator is diagonalized by a 2$`\times `$2 complex orthogonal matrix which can be parametrized as usual by one complex variable,
$$๐ช=\left(\begin{array}{cc}\mathrm{cos}\chi & \mathrm{sin}\chi \\ \mathrm{sin}\chi & \mathrm{cos}\chi \end{array}\right),$$
(38)
where in this way we introduce the $`NN^{}`$ mixing angle $`\chi `$.
Furthermore, since we use the same Feynman rules for the nucleon and $`N^{}`$, their dressed vertices are equal up to the coupling constants. Therefore for the $`NN^{}`$ self-energy matrix one can write
$$\left(\begin{array}{cc}\Sigma _{NN}(P)& \Sigma _{NN^{}}(P)\\ \Sigma _{NN^{}}(P)& \Sigma _{N^{}N^{}}(P)\end{array}\right)=\left(\begin{array}{cc}g_{\pi NN}^2& g_{\pi NN}g_{\pi NN^{}}\\ g_{\pi NN}g_{\pi NN^{}}& g_{\pi NN^{}}^2\end{array}\right)\Sigma (P),$$
(39)
while for the vertex
$$\left(\begin{array}{c}\Gamma _N^{}\\ \Gamma _N^{}^{}\end{array}\right)=\left(\begin{array}{c}g_{\pi NN}\\ g_{\pi NN^{}}\end{array}\right)\Gamma ^{}.$$
(40)
The propagator is then diagonalized by the orthogonal transformation (38) with
$$\chi (P)=\frac{1}{2}\mathrm{arctg}\left\{2\left(\frac{g_{\pi NN}}{g_{\pi NN^{}}}\frac{g_{\pi NN^{}}}{g_{\pi NN}}\frac{m_N^{}m_N}{g_{\pi NN}g_{\pi NN^{}}\Sigma (P)}\right)^1\right\}.$$
(41)
The corresponding eigenvalues are clearly
$`S_N(P)`$ $`=`$ $`[P/m_N\Sigma _N(P)iฯต]^1,`$ (42)
$`S_N^{}(P)`$ $`=`$ $`[P/m_N^{}\Sigma _N^{}(P)iฯต]^1,`$ (43)
where
$`\Sigma _N`$ $`=`$ $`\left(g_{\pi NN}\mathrm{cos}\chi +g_{\pi NN^{}}\mathrm{sin}\chi \right)^2\Sigma +(m_N^{}m_N)\mathrm{sin}^2\chi ,`$ (44)
$`\Sigma _N^{}`$ $`=`$ $`\left(g_{\pi NN^{}}\mathrm{cos}\chi g_{\pi NN}\mathrm{sin}\chi \right)^2\Sigma (m_N^{}m_N)\mathrm{sin}^2\chi .`$ (45)
The vertices are rotated according to
$`\Gamma _N^{}=(g_{\pi NN}\mathrm{cos}\chi +g_{\pi NN^{}}\mathrm{sin}\chi )\Gamma ^{},`$ (46)
$`\Gamma _N^{}^{}=(g_{\pi NN^{}}\mathrm{cos}\chi g_{\pi NN}\mathrm{sin}\chi )\Gamma ^{}.`$ (47)
### B Self-energy
Let us consider the mass renormalization using the counter-term method. The counter terms can be read off directly from the free Lagrangian. For the spin-1/2 case, for instance, they are given by $`Z_2(m_0m)+(1Z_2)(P/m)`$, where $`m_0`$ is the bare mass, and $`Z_2`$ is the field renormalization constant. The renormalized spin-1/2 baryon propagator is defined as
$$๐(P/)=[P/m\Sigma ^{ren}(P/)i\epsilon ]^1,$$
(48)
where $`\Sigma ^{ren}(P/)=\Sigma (P/)Z_2(m_0m)(1Z_2)(P/m)`$.
In the CMS, $`P=(P_0,\stackrel{}{0})`$, the self-energy can be written as
$$\Sigma ^{ren}(P_0)=\Sigma _+^{ren}(P_0)\gamma _++\Sigma _{}^{ren}(P_0)\gamma _{},$$
(49)
and a similar decomposition holds for the propagator,
$$๐(P_0)=๐^{(+)}(P_0)\gamma _++๐^{()}(P_0)\gamma _{}$$
(50)
where $`\gamma _\pm =(I\pm \gamma _0),`$ and
$$\begin{array}{c}๐^{(+)}(P_0)=\left[P_0m\Sigma _+^{ren}(P_0)i\epsilon \right]^1,\hfill \\ ๐^{()}(P_0)=\left[P_0+m+\Sigma _{}^{ren}(P_0)+i\epsilon \right]^1.\hfill \end{array}$$
(51)
Obviously, $`\gamma _+`$ and $`\gamma _{}`$ act as the projection operators onto the positive and negative energy-states, hence $`๐^{(+)}`$ corresponds to the positive and $`๐^{()}`$ to the negative energy-state propagations.
The renormalization condition at the pole position is given by
$$\begin{array}{c}(P_0m)๐^{(+)}(P_0)|_{P_0=m}=1,\hfill \\ (P_0+m)๐^{()}(P_0)|_{P_0=m}=1.\hfill \end{array}$$
(52)
Expanding $`\Sigma _+(P_0)`$ near $`P_0=m`$, and $`\Sigma _{}(P_0)`$ near $`P_0=m,`$ we find that the renormalization condition requires
$`Z_2(m_0m)`$ $`=`$ $`\Sigma _+(m)=\Sigma _{}(m),`$ (53)
$`1Z_2`$ $`=`$ $`{\displaystyle \frac{\Sigma _+(P_0)}{P_0}}|_{P_0=m}={\displaystyle \frac{\Sigma _{}(P_0)}{P_0}}|_{P_0=m}.`$ (54)
As emphasized earlier , the above described procedure breaks down if the self-energy is computed using a quasipotential formulation which violates charge conjugation symmetry, since in that case $`\Sigma _+(P_0)\Sigma _{}(P_0)`$. The self-energy of the spin-3/2 baryons can be renormalized similarly, since the spin-3/2 baryon contribution to the spin-3/2 partial-waves can always be factorized into vertices and a spin-1/2 propagator, see Eq. (D13).
### C The renormalized vertex and the amplitudes
In the adopted renormalization scheme we require (i) the (real part of) renormalized baryon self-energy $`\Sigma _\pm ^{ren}(P_0)`$ and its first derivative vanish at the pole position $`P_0=\pm m`$; (ii) the (real part of) renormalized vertex $`\pi N`$baryon vertex is equal to the undressed vertex at the renormalization scale $`\mu `$ defined as the point where all three particles are on the mass-shell, $`\mu `$: $`k^2=m_\pi ^2,p^2=m_N^2,P^2=m^2`$.
For the vertex we use the multiplicative renormalization since it maintains unitarity in a simple way. The renormalized vertex is thus defined as
$$๐ช(p;P)=Z_1\Gamma ^{}(p;P),$$
(55)
where $`Z_1`$ is the coupling constant renormalization factor which is readily determined from condition (ii):
$$Z_1=๐ช(\mu )/\Gamma ^{}(\mu ).$$
(56)
In the case of the $`NN^{}`$-mixing we renormalize the (scalar) function $`\mathrm{\Sigma }`$ in Eq. (39) at the point associated with the nucleon. This procedure clearly yields the proper physical nucleon mass pole in the corresponding baryon propagator. Adopting this subtraction procedure the $`N^{}`$ mass and coupling constant at the nucleon mass position can then be extracted. In the various tables the values of these parameters found in the fits are quoted.
After the partial-wave decomposition, the renormalized solution of the ET equation for a given isospin $`I`$ and total spin $`J`$ and parity $`r`$ reads as follows (for brevity the external momenta are omitted):
$`T_r^{J\rho ^{}\rho }`$ $`=`$ $`{\displaystyle \underset{\pi }{}}๐ช_r^\rho ^{}๐^{(r\eta _B)}๐ช_r^\rho \delta _{J_BJ}\delta _{I_BI}+T_{u,r}^{J\rho ^{}\rho },`$ (57)
$`T_{u,r}^{J\rho ^{}\rho }`$ $`=`$ $`V_{u,r}^{J\rho ^{}\rho }+{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dqq^2{\displaystyle \underset{\rho ^{\prime \prime }}{}}V_{u,r}^{J\rho ^{}\rho ^{\prime \prime }}(q)G_{ET}^{(\rho ^{\prime \prime })}(q)T_{u,r}^{J\rho ^{\prime \prime }\rho }(q),`$ (58)
$`๐ช_r^\rho `$ $`=`$ $`Z_1\left(\Gamma _r^\rho +{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dqq^2{\displaystyle \underset{\rho ^{}}{}}\Gamma _r^\rho ^{}(q)G_{ET}^{(\rho ^{})}(q)T_{u,r}^{J\rho ^{}\rho }(q)\right),`$ (59)
$`\Sigma _r`$ $`=`$ $`Z_1{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}dqq^2{\displaystyle \underset{\rho }{}}\Gamma _r^\rho (q)G_{ET}^{(\rho )}(q)๐ช_r^\rho (q),`$ (60)
where $`J_B,I_B`$ and $`\eta _B`$ are the baryon spin, isospin and parity, respectively. The renormalized propagator is
$$๐^{(\pm )}(P_0)=\left[\pm P_0m\Sigma _\pm ^{ren}(P_0)+i\epsilon \right]^1,$$
(61)
where the renormalized self-energy is given in terms of $`\Sigma `$ of Eq. (60) \[with Eq. (45) in the case of $`NN^{}`$-mixing\] as follows:
$`\Sigma _\pm ^{ren}(P_0)`$ $`=`$ $`\Sigma _\pm (P_0)\Sigma _\pm (\pm m)`$ (62)
$``$ $`(P_0m){\displaystyle \frac{\Sigma _\pm (P_0)}{P_0}}|_{P_0=\pm m}.`$ (63)
## V Results
Having described the equation for the off-shell $`\pi N`$ amplitudes, its renormalization, and the driving force, we now turn to discussing the outcome of such modeling. For this let us first give an explicit relation between the on-shell amplitudes $`T_r^{J++}`$ and the phase parameters.
We introduce the standard on-shell $`\pi N`$ amplitudes $`f_{l\pm }`$, where $`l=J\frac{1}{2}r`$ is the angular momentum and $`r`$ is the parity of the state. In the normalization according to
$$f_{l\pm }=\frac{\eta _{l\pm }e^{2i\delta _{l\pm }}1}{2i},$$
(64)
where $`\delta _{l\pm }`$ is the phase shift and $`\eta _{l\pm }`$ is the inelasticity, we can identify
$$f_{l\pm }=\widehat{q}\frac{s+m_N^2m_\pi ^2}{2s}T_r^{J++}(\widehat{E},\widehat{q},\widehat{E},\widehat{q};P_0),$$
(65)
where
$$\widehat{q}=\sqrt{[s(m_Nm_\pi )^2][s(m_N+m_\pi )^2]/4s},$$
(66)
and $`\widehat{E}=\sqrt{m_N^2+\widehat{q}^2}`$ is the energy of the nucleon in the CMS.
At very low energies the partial-wave amplitudes are dominated by the threshold behavior: $`\widehat{q}^{2l+1}`$, and their real and imaginary parts are related by elastic unitarity. Therefore, it is sometimes more useful to study
$$_{lJ}(\widehat{q}^2)=\left(\frac{m_\pi }{\widehat{q}}\right)^{2l+1}\text{Re}f_{l\pm },$$
(67)
instead of $`f_{l\pm }`$ itself or the phase shifts. Note that, at $`\widehat{q}^2=0`$, $`_{lJ}`$ is equal to the corresponding scattering length defined asWe shall commonly refer to this quantity as to scattering length, even though for $`P`$ and higher waves it is properly called scattering volume because of the dimension.
$$a_{lJ}\underset{\widehat{q}0}{lim}[\widehat{q}^{2l1}f_{l\pm }(\widehat{q})].$$
(68)
The $`\pi N`$ effective-range parameters $`b_{lJ}`$ can also be determined in terms of $``$:
$$b_{lJ}=\frac{}{\widehat{q}^2}_{lJ}(\widehat{q}^2)|_{\widehat{q}=0}.$$
(69)
Instead of using this formula, we will be presenting the plot of $``$ as a function of $`\widehat{q}^2`$. The slope of these โeffective-range plotsโ at small $`\widehat{q}`$ indicates the values for the effective-range parameters. In the partial waves that support a resonance (e.g., $`P_{11}`$, $`P_{33}`$) it is more appropriate to study another effective-range expansion:
$$\widehat{q}^{2l+1}\mathrm{cot}\delta _{l\pm }=\frac{1}{a_{lJ}}+\frac{1}{2}r_{lJ}\widehat{q}^2+\mathrm{},$$
(70)
nontheless, we only will address threshold parameters $`a`$ and $`b`$ for all partial waves.
### A The $`K`$-matrix approximation
In this subsection we focus on the K-matrix approximation to the full scattering problem. Results of the full model are discussed in the next subsection. In the K-matrix approximation $`\pi N`$ amplitude is given by the lowest order $`K`$-matrix expression:
$$f_{l\pm }=\frac{K_{l\pm }}{1iK_{l\pm }},$$
(71)
where $`K_{l\pm }=\widehat{q}\alpha (s)V_\pm ^J,l=J\frac{1}{2}`$, and $`V_r^J`$ is the partial-wave potential \[obtained from the potential by means of Eqs. (13) and (B12)\]. Equation (71) clearly satisfies elastic unitarity, but the principal value of the loop integrals is neglected.
This approximation is considered to be good, at least at low energy, and has been frequently used, see e.g. for most recent applications to the $`\pi N`$ scattering. At low energies, indeed, the soft-pion theorems dictate that the Born graphs dominate, implying for the potential modelling that the rescattering effects should be relatively small. When the latter is true, considering the K-matrix approximation may allow us to make a preliminary adjustment of some model parameters without going to the full calculations. We in particular would like to examine the effect of using different $`\pi N\mathrm{\Delta }`$ couplings in the $`\mathrm{\Delta }`$ exchange contribution.
As is well known, the S-wave scattering lengths are well reproduced by the Born-level nucleon and $`\rho `$-meson exchanges alone . Taking $`g_{\pi NN}^2/4\pi =13.6`$ and $`g_\rho ^2/4\pi =3.0`$, we plot in Fig. 6 the contribution of these three graphs ($`s`$\- plus $`u`$-channel nucleon exchange plus $`t`$-channel $`\rho `$ exchange) to $`_{lJ}`$ for all the $`S`$ and $`P`$ partial waves. The figure shows that the $`S`$-wave scattering lengths are indeed reproduced. However, the energy dependence of the $`S`$ waves is not well described. In fact, the slopes of the effective-range plots have a wrong sign. Furthermore, the $`P`$ waves are not reproduced at all. Similar picture occurs if the contact term is used instead of the $`\rho `$ exchange.
Looking at the ratio of the $`S`$-wave lengths it is certainly plausible that they should be dominated by some isovector contribution, Experimentally the ratio $`a_{S11}/a_{S31}1.75`$, which is close to the ratio of the isospin factors for an isovector meson contribution. such as the $`\rho `$-meson exchange. Therefore, it would be interesting to find a simple mechanism which accounts for both the $`P`$ waves and the energy behavior of the $`S`$ waves, and, at the same time, does not affect the $`S`$-wave scattering length. Since $`P_{33}`$ has the largest discrepancy, we study first the effect of the $`\mathrm{\Delta }`$-isobar exchange.
In Fig. 7 we show the calculations performed with the two different choices of the off-shell parameter:
$$\begin{array}{cc}z=\frac{1}{4}\hfill & \text{(Peccei choice }\text{[28]}\text{)},\hfill \\ z=\frac{1}{2}\hfill & \text{(NEK choice }\text{[29]}\text{)},\hfill \end{array}$$
and $`f_{\pi N\mathrm{\Delta }}^2/4\pi =0.36`$. The nucleon and $`\rho `$ coupling constants are kept the same as in the previous calculation, with $`\kappa _\rho =3.7`$.
One can see that, as far as $`P_{33}`$ is concerned, the $`\mathrm{\Delta }`$ contribution is very plausible for both choices of $`z`$. The large difference between the choices clearly shows up in the spin-1/2 partial waves, where the $`\mathrm{\Delta }`$ exchange produces a significant background contribution controlled by the off-shell parameter.
In particular, Pecceiโs choice affects substantially the $`S`$ waves, and hence spoils the scenario where those are dominated by the $`\rho `$-exchange. The $`P`$ waves look much better, which could be due to the remarkable fact that, for this particular choice, the $`s`$-channel $`\mathrm{\Delta }`$-exchange graph gives only a tiny contribution to the spin-1/2 $`P`$ waves. The NEK choice, in contrast, does not affect the $`S`$ waves at threshold and gives a large effect in the $`P`$ waves.
Clearly, Pecceiโs choice could be favorable phenomenologically as long as the missing strength in the $`S`$ waves is somehow explained; for instance, by an isoscalar meson exchange. We believe such a scenario is realized in most of the models which use coupling (25) and describe the scattering lengths correctly. However, apparently it is not possible to describe simultaneously the $`S`$\- and $`P`$-wave scattering lengths in the tree-level model with only the $`\rho `$, $`N`$ and $`\mathrm{\Delta }`$ exchanges.
The GI $`\pi N\mathrm{\Delta }`$ coupling Eq. (26), in combination with the usual Rarita-Schwinger propagator, leads to the $`\mathrm{\Delta }`$-exchange amplitude with only spin-3/2 contributions (omitting the isospin-dependent factor):
$`V_{\mathrm{\Delta },s\mathrm{exch}}={\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi ^2}}\left({\displaystyle \frac{s}{m_\mathrm{\Delta }^2}}\right){\displaystyle \frac{P/+m_\mathrm{\Delta }}{sm_\mathrm{\Delta }^2}}๐ซ_{\alpha \beta }^{3/2}(P)k_{}^{}{}_{}{}^{\alpha }k^\beta ,`$ (72)
$`V_{\mathrm{\Delta },u\mathrm{exch}}={\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi ^2}}\left({\displaystyle \frac{u}{m_\mathrm{\Delta }^2}}\right){\displaystyle \frac{p/_u+m_\mathrm{\Delta }}{um_\mathrm{\Delta }^2}}๐ซ_{\alpha \beta }^{3/2}(p_u)k^\alpha k_{}^{}{}_{}{}^{\beta },`$ (73)
where $`p_u=Pkk^{}`$, and
$$๐ซ_{\alpha \beta }^{3/2}=\mathrm{g}_{\alpha \beta }\frac{1}{3}\gamma _\alpha \gamma _\beta \frac{1}{3p^2}(p/\gamma _\alpha p_\beta +p_\alpha \gamma _\beta p/)$$
is the spin-3/2 projection operator.
A calculation using this amplitude is shown in Fig. 7, in comparison with analogous calculations using the conventional coupling with Peccei and NEK choices of the off-shell parameter. First of all we remark that the $`\mathrm{\Delta }`$ contribution to the spin-1/2 partial-waves comes from the $`u`$-graph only, and not from the spin-1/2 components. From the figure we can see that the GI and NEK coupling produce similar contributions to the $`S`$-waves, but largely different results in the spin-1/2 $`P`$-waves. The resonant $`P_{33}`$ wave comes out very much alike for both GI and conventional couplings, the main difference being $`(s/m_\mathrm{\Delta }^2)`$ factor in front of the GI amplitude. Hence, at threshold the GI result is factor of $`(m_N+m_\pi )^2/m_\mathrm{\Delta }^20.76`$ smaller than the conventional result (modulo small contributions from the $`u`$-channel graph, which for instance are responsible for the difference between the Peccei and NEK choice in the $`P_{33}`$ wave). Despite that, after a readjustment of parameters the GI invariant coupling usually gives a better description of the scattering lengths, see calculations presented by Tables I and II (note that here the NEK choice has been used with $`\kappa _\rho =0`$, as is suggested in the original paper and indeed gives a better fit than with $`\kappa _\rho =3.7`$). The problem with the wrong energy behavior of the $`S`$ waves, however, applies to all these calculations. To correct for this a scalar $`\sigma `$-meson exchange is needed in the $`\pi N`$ force.
Inclusion of the $`\sigma `$-exchange allows us to fit the scattering lengths to practically arbitrary accuracy, independently of whether we use the conventional (model A) or the GI coupling (model B), see Tables III and IV. Since we have fixed $`c_\sigma =1`$, and thus do not allow the $`\sigma `$ to affect the $`S`$-waves, the best fit of the off-shell parameter give the NEK value which also has vanishing $`S`$-wave contribution. The $`S`$-wave lengths are therefore explained in both models A and B in exactly the same way: by $`\rho `$ and nucleon exchanges alone. The small difference between the NEK and GI coupling in $`P_{33}`$ (see Fig. 7) is apparently compensated by the difference in $`f_{\pi N\mathrm{\Delta }}`$. The large differences in the other three partial-waves has mostly been removed by taking a different value for $`\kappa _\rho `$. This example shows that if the potential is general enough, the spin-1/2 backgrounds of the $`\mathrm{\Delta }`$ can possibly be reshuffled into other contributions.
### B Full calculations
We solve the ET equation, Eq. (14), by Padรฉ approximants following the procedure described in Refs. . Writing the equation as
$$T=V+VGT,$$
(74)
we begin by performing several iterations of the potential, and hence find the first few terms in the expansion of the amplitude:
$$T=\underset{n=0}{\overset{\mathrm{}}{}}T^{(n)}=V+\underset{n=1}{\overset{\mathrm{}}{}}V(GV)^n.$$
(75)
The solution is then sought in the form of the Padรฉ approximant. For the equal-time equation with the model potential the solution accurately converges by performing just six iterations.
We have fitted the on-shell solution to the KH80 and SM95 $`\pi N`$ scattering partial-wave analyses, in the region from the threshold up to 600 MeV pion kinetic energy in the lab. The resulting fit is shown by the solid lines in Fig. 8. The determined coupling constants and masses are given in Table V.
The dashed line in the $`S_{11}`$ phase-shift of Fig. 8 indicates the calculation without the $`s`$-channel $`S_{11}`$ resonance graph. This graph contributes also to the $`P_{11}`$ wave but calculation with or without it produce practically identical results. The $`S_{11}`$ resonance pole is thus relevant only for the $`S_{11}`$ wave above 400 MeV.
Up to 350 MeV the agreement of the model with the partial-wave data is very good as can be seen in Fig. 9. At energies exceeding 350 MeV inelastic channels become important. Since we have not considered any inelastic mechanisms, some discrepancies seen in Fig. 8 at higher energies are not surprising.
In fitting we have paid careful attention to the low-energy behavior, particularly to the correct shape of the โeffective-range plotโ, $`_{l\pm }(\widehat{q}^2)`$, defined in Eq. (67). The model description of these plots is shown in Fig. 10. The scattering lengths can be read off these plots at $`\widehat{q}=0`$. The $`\chi `$-square value for the scattering lengths with respect to the SM95 analysis is $`1.4`$.
To give a feeling about the size of the rescatterings in the model, the dashed lines in the figures indicate the results of the calculation where the principal part of the rescattering integrals is neglected, i.e., the $`K`$-matrix approximation. Unlike in the $`K`$-matrix calculations of the previous subsection, here the form factors are included, and the same set of parameters is used as in the full calculation.
The large difference between the full and the $`K`$-matrix calculation in the waves where the baryon pole contributions are present indicates, therefore, that the โdressingโ and renormalization of the pole contributions has a significant effect. The effect of the rescatterings in the nonresonant waves, such as $`S_{31}`$, $`P_{31}`$, etc., which have contribution only from the $`T_u`$ term, is smaller. We also observed that the โattractive wavesโ (which have positive scattering length) receive comparably large and positive rescattering contributions at threshold.
For the resonant waves the $`T_u`$ contribution may lead to significant shifts of the resonance position. For instance, the $`\mathrm{\Delta }`$-pole in the P<sub>33</sub> is located at $`\sqrt{s}=m_\mathrm{\Delta }=1.252`$ GeV, while inclusion of $`T_u`$ leads to the physical $`P_{33}`$ wave which resonates at $`\sqrt{s}=1.232`$ GeV. In the $`K`$-matrix approximation the resonance always occurs at the pole position. To describe $`P_{33}`$ in this approximation one is to use $`m_\mathrm{\Delta }=`$1.232 GeV.
By readjusting the parameters we are able to reproduce the phase-shifts in the $`K`$-matrix approximation up to 350 MeV, see Fig. 11. The values of the parameters are given in Table VI. Note, however, that using the full model we could achieve the fit of a better quality and up to higher energies.
The ($`s`$-channel) baryon pole terms are modified by the vertex and self-energy corrections. As has been seen from comparing the full and $`K`$-matrix calculation, such a โdressingโ may have appreciable effects, especially in the resonance $`P_{11}`$ and $`P_{33}`$ waves. In Fig. 12 we plot the real and imaginary part of the nucleon and the $`\mathrm{\Delta }`$ isobar self-energies. From the figure we can see that the energy dependence is indeed significant. The same is observed for the $`NN^{}`$ mixing angle plotted in Fig. 13. The renormalization of the pole terms produces the values of the renormalization constants given in Table VII.
The vertex corrections are studied using the dynamical form factor defined as follows:
$$F^\rho (q^2,s)=๐ช^\rho (q^2,s)/\Gamma ^\rho (q^2,s),$$
(76)
where $`\Gamma ^\rho `$ is the undressed off-shell vertex, and $`๐ช^\rho `$ is the renormalized off-shell vertex, see Eq. (59). Note that the coupling constants and the cutoff form factors are cancelling out in the expression (76), since they are the same for both of the vertices. The dynamical form factors are thus equal to unity at the renormalization point.
The model prediction for $`\pi NN`$, $`\pi N\mathrm{\Delta }`$ and $`\pi ND_{13}`$ form factors is given in Fig. 14 as a function of the off-shell 3-momentum $`\stackrel{}{q}^2`$ for $`\sqrt{s}=m_N+m_\pi `$, and in Fig. 15 as the function of energy for the on-shell situation, $`|\stackrel{}{q}|=\widehat{q}`$. According to these figures the rescatterings have much larger effect on the energy dependence then on the off-shell behavior of the $`\pi N`$ state.
Recently, the $`\pi NN`$ form factor has been studied by Saito and Afnan , and by Schรผltz, Haidenbauer and Holinde in a similar modelling. In the latter work the $`\pi N\mathrm{\Delta }`$ form factor has been studied as well. To compare with the results presented in , we need to multiply our dynamical form factor by the off-shell form factor given in Fig. 3 (solid line). We then can see that the resulting $`\pi NN`$ form factor of our model agrees in the main features with that of model . It is therefore less soft than the form factor found in Ref. . This allows the rescattering contributions to play a bigger role.
## VI Discussion and conclusion
Throughout the calculation we have been fixing the $`\pi NN`$ coupling constant to the value advised by the Nijmegen group : $`f_{\pi NN}^2/4\pi =0.0757`$. Our fits were not very sensitive to the increase of the coupling towards more traditional value of 0.078.
The value of $`g_\rho `$ comes out close to the one inferred by the $`\rho `$-meson decay width: $`g_\rho ^2/4\pi 2.8`$. It is also consistent with the KSFR relation (Sec. IV A), which gives 2.78 in its first form, and 2.94 in the second form. The small value of $`\kappa _\rho `$ has been mainly dictated by the simultaneous fit to $`S_{31}`$ and $`P_{31}`$ waves at the higher energy scale. At low energies the change in $`\kappa _\rho `$ affects mostly the $`P_{11}`$ channel, as can be seen from Fig. 10 where the dash-dotted line represents the calculation with the vector meson dominance value: $`\kappa _\rho =3.7`$.
Comparing Tables V and VI we see that, depending on whether the rescattering contributions are included or not, very different values of the $`\mathrm{\Delta }`$-isobar masses and coupling strengths are needed to obtain correct phase shifts. This indicates how the dynamical component due to $`\pi N`$ loops may play a significant role in the generation of the observed $`\mathrm{\Delta }`$(1232) resonance. That is in addition to the โelementaryโ component due to the formation of the three-quark state.
In comparing with other relativistic models we can comment that our major difference with the model of Pearce and Jennings resides in the $`\mathrm{\Delta }`$-isobar and $`\sigma `$-meson contributions. For the $`\mathrm{\Delta }`$, they use the conventional coupling, and for the $`\sigma \pi \pi `$ coupling they use Eq. (21) with the opposite sign and without the additional $`c_\sigma `$ term Eq. (24). The $`\sigma `$ contribution is thus attractive in their case and can account for the discrepancy in the $`S`$-waves which appears due to the $`\mathrm{\Delta }`$ background. Although it probably gives rise to problems at higher energies, which is indicated by the very soft cutoff form factor for the $`\sigma `$, with cutoff mass $`\Lambda _\sigma =0.5`$ GeV, used in .
Gross and Surya made a static approximation of all the $`t`$\- and $`u`$\- channel graphs. Their approximation leads to separability of the potential and hence the complexity of solving the integral equation numerically is avoided. It also simplifies considerably the subsequent photoproduction analysis, since the meson- and isobar-exchange currents have a form of contact terms. On the other hand, the spin-3/2 (and higher) waves, such as $`P_{33}`$ and $`D_{13}`$, receive contributions only from the $`s`$-channel exchanges of corresponding resonances, which in particular leads to overestimates of the resonance coupling parameters. Also, the static (zero-range) approximation of the $`u`$-exchange potential can be justified only at low energy, because in principle the range of such a potential rapidly increases from $`1/\sqrt{2mm_\pi }`$ (where $`m`$ is the exchanged particle mass) at threshold till $`1/m_\pi `$ at high energy. We observed that already at 100 MeV pion kinetic energy the solution of the integral equation for the static or exact $`u`$-exchange potential differ significantly. In contrast, the range of the $`t`$-exchange potential is determined only by the exchange mass, and if that mass is heavy enough the static approximation may be applicable. Another difference comes from the fact that Gross and Surya use the admixture of pseudoscalar and pseudovector coupling for the $`\pi NN`$ vertex. Consequent differences, motivated by the consistency with the soft-pion limit, appear in the form of the $`\rho `$ and $`\sigma `$ couplings.
Lahiff and Afnan do not include resonances beyond the $`\mathrm{\Delta }`$(1232) but apart from that they use an interaction, which is very similar to ours. They have as well compared the conventional versus gauge-invariant $`\pi N\mathrm{\Delta }`$ coupling. However, in contrast to us, they preferred the former one, particularly because the spin-1/2 background adds a significant attraction in the $`P_{11}`$ channel, which helps to simulate partially the Roper-resonance behavior. We include this resonance explicitly and hence the spin-1/2 background is not necessary for fitting $`P_{11}`$ phase shift.
Including the $`\mathrm{\Delta }`$ and $`D_{13}`$ resonance in the Lagrangian via the relativistic spin-3/2 fields, we have used couplings which do not involve the spin-1/2 sector of the Rarita-Schwinger theory, and therefore no โspin-1/2 backgroundsโ associated with these particles appear in the $`S`$-matrix. We do not need these backgrounds to obtain a proper description of the data. On a simple example (see Tables III and IV) we have seen that, even when these backgrounds may sometimes seem relevant for the description, their role can be taken by other mechanisms, hopefully with more sensible physical interpretation.
Although the presented model improves in some aspects previous relativistic analyses based on the potential approach, the difficulty of controlling the chiral symmetry constraint is present here as well. We do begin with a Lagrangian, and thus the driving force, consistent with chiral symmetry, but this consistency can in principle be spoiled by rescatterings, particularly because of the loss of crossing symmetry. However, numerical checks (see, e.g., Ref. ) indicate that the amount of violation of the soft-pion theorems is usually negligible in the potential modeling. Especially if to take into account that the prime objection of such models is to describe the $`\pi N`$ physics at intermediate energies where unitarity aspects take the leading role. It would, of course, be anyhow important to build the chiral constraint more precisely into the nonperturbative $`\pi N`$ models.
In conclusion, we have obtained a description of the $`\pi N`$ force in a relativistic dynamical model based on the covariant equal-time (quasipotential) equation for a hadron-exchange potential. The good quality of our fits in the region up to 600 MeV pion kinetic energy suggests the used force and the relativistic approach may be considered reasonable, even though this model still lacks inelastic mechanisms which can become important in $`S_{11}`$ and $`P_{11}`$ channels above 400 MeV. It is clearly of interest to study the model predictions in other processes, such as pion photoproduction and Compton scattering in the $`\pi N`$ system.
###### Acknowledgements.
This work was partially financially supported by de Stichting voor Fundamenteel Onderzoek der Materie (FOM), which is sponsored by Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO).
## A Conventions
Metric $`\mathrm{g}_{\mu \nu }=\mathrm{diag}(+1,1,1,1)`$; Levy-Cevita symbols: $`\epsilon ^{0123}=1`$, $`\epsilon ^{0ijk}=\epsilon ^{ijk}`$.
Pauli spinors: in the direction $`\theta ,\phi `$ are given by
$$\chi _\lambda (\theta ,\phi )=\underset{\lambda ^{}=1/2}{\overset{1/2}{}}d_{\lambda \lambda ^{}}^{1/2}(\theta )e^{i(\lambda \lambda ^{})\phi }\chi _\lambda ^{}(0),$$
where $`\chi _{1/2}^{}(0)=(1,0)`$, $`\chi _{1/2}^{}(0)=(0,1)`$, and $`d_{\lambda \lambda ^{}}^J(\theta )`$ are the Wigner $`d`$-functions.
Positive- and negative-energy helicity spinors:
$`u_\lambda ^{(+)}(\stackrel{}{p})`$ $`=`$ $`\left[\begin{array}{c}N_+\\ 2\lambda N_{}\end{array}\right]\chi _\lambda (\theta ,\phi ),`$ (A3)
$`u_\lambda ^{()}(\stackrel{}{p})`$ $`=`$ $`\left[\begin{array}{c}2\lambda N_{}\\ N_+\end{array}\right]\chi _\lambda (\theta ,\phi ),`$ (A6)
where $`N_\pm =\sqrt{(E_p\pm m)/2E_p}`$, $`E_p=\sqrt{m^2+\mathrm{p}^2}`$, and $`\mathrm{p}`$, $`\theta `$, $`\phi `$ are the spherical coordinates of $`\stackrel{}{p}`$. The helicity spinors satisfy the following orthogonality and completeness conditions:
$`u_\lambda ^\rho (\stackrel{}{p})u_\lambda ^{}^\rho ^{}(\stackrel{}{p})=\delta _{\rho \rho ^{}}\delta _{\lambda \lambda ^{}},`$ (A7)
$`{\displaystyle \underset{\rho =\pm }{}}{\displaystyle \underset{\lambda }{}}u_\lambda ^\rho (\stackrel{}{p})u_\lambda ^\rho (\stackrel{}{p})=1.`$ (A8)
## B Partial-wave off-shell $`\pi N`$ amplitudes
The partial-wave reduction is done in the CMS, where the total four-momentum $`P=(P_0,\stackrel{}{0})`$. Using the orthonormality and completeness of the nucleon helicity spinors (see Appendix A), we write the BS equation for the helicity amplitudes:
$$T_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p)=V_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p)+i\underset{\rho ^{\prime \prime }\lambda ^{\prime \prime }}{}\frac{\mathrm{d}^4q}{4\pi ^3}V_{\lambda ^{}\lambda ^{\prime \prime }}^{\rho ^{}\rho ^{\prime \prime }}(p^{},q)G^{(\rho ^{\prime \prime })}(q)T_{\lambda ^{\prime \prime }\lambda }^{\rho ^{\prime \prime }\rho }(q,p),$$
(B1)
where we have assumed that the $`\pi N`$ propagator is diagonal in the helicity basis, i.e.,
$$\overline{u}_\lambda ^{}^\rho ^{}(\stackrel{}{q})\gamma _0G(q)\gamma _0u_{\lambda ^{\prime \prime }}^{\rho ^{\prime \prime }}(\stackrel{}{q})=\delta _{\lambda ^{}\lambda ^{\prime \prime }}\delta _{\rho ^{}\rho ^{\prime \prime }}G^{(\rho ^{\prime \prime })}(q),$$
(B2)
which is true in the CMS. Eq. (B1) yields (omitting the momenta)
$$T_{\lambda ^{}\lambda }^{J\rho ^{}\rho }=V_{\lambda ^{}\lambda }^{J\rho ^{}\rho }+\frac{i}{\pi ^2}\underset{\mathrm{}}{\overset{\mathrm{}}{}}dq_0\underset{0}{\overset{\mathrm{}}{}}dqq^2\underset{\rho ^{\prime \prime }\lambda ^{\prime \prime }}{}G^{(\rho ^{\prime \prime })}V_{\lambda ^{}\lambda ^{\prime \prime }}^{J\rho ^{}\rho ^{\prime \prime }}T_{\lambda ^{\prime \prime }\lambda }^{J\rho ^{\prime \prime }\rho },$$
(B3)
where
$`X_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p;P_0)={\displaystyle \underset{J=1/2}{\overset{\mathrm{}}{}}}(J+\frac{1}{2})X_{\lambda ^{}\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0)d_{\lambda ^{}\lambda }^J(\theta ),`$ (B4)
$`X_{\lambda ^{}\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0)={\displaystyle _1^1}\mathrm{d}(\mathrm{cos}\theta )X_{\lambda ^{}\lambda }^{\rho ^{}\rho }(p^{},p;P_0)d_{\lambda ^{}\lambda }^J(\theta ),`$ (B5)
with $`X=V`$ or $`T`$. We have chosen the 3-vectors $`\stackrel{}{p}`$ and $`\stackrel{}{p}^{}`$ to lie in the $`XZ`$-plane (hence $`\phi =\phi ^{}=0`$), and $`\theta `$ is the center-of-mass scattering angle.
Parity conservation infers the following symmetry for the partial-wave helicity amplitudes:
$$T_{\lambda ^{}\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0)=\rho ^{}\rho T_{\lambda ^{},\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0).$$
(B6)
These relations reduce the number of independent amplitudes from sixteen to eight. Time-reversal invariance implies
$$T_{\lambda ^{}\lambda }^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0)=4\lambda ^{}\lambda T_{\lambda \lambda ^{}}^{J\rho ^{}\rho }(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0),$$
(B7)
which obviously does not give any new relations.
It is convenient to introduce the partial-wave state with definite parity $`r`$ in terms of the partial-wave helicity state :
$$|J,r,\rho =\frac{|J,\rho ,\lambda +r\rho |J,\rho ,\lambda }{\sqrt{2}}.$$
(B8)
The BS equation for the parity-conserving amplitudes takes the form
$$T_r^{J\rho ^{}\rho }=V_r^{J\rho ^{}\rho }+\frac{i}{\pi ^2}\underset{\mathrm{}}{\overset{\mathrm{}}{}}dq_0\underset{0}{\overset{\mathrm{}}{}}dqq^2\underset{\rho ^{\prime \prime }}{}G^{(\rho ^{\prime \prime })}V_r^{J\rho ^{}\rho ^{\prime \prime }}T_r^{J\rho ^{\prime \prime }\rho },$$
(B9)
where $`r`$ denotes parity. Note the simplification: Eq. (B9) has two less coupled channels than Eq. (B3). The partial-wave decomposition of the ET equation (10) proceeds in the same way. In addition we use in this case the fact that in the CMS $`V`$ and $`T`$ are independent of $`q_0`$. The resulting equations are given by Eq. (14).
The amplitudes (13) are related in a simple way to the parity-conserving partial-wave amplitudes of Jacob and Wick . Starting from representation Eq. (13), we first reduce the spinor algebra to the subspace of Pauli spinors $`\chi _\lambda (\theta )`$ and use
$$\chi _\lambda ^{}^{}(\theta ^{})\chi _\lambda (\theta )=d_{\lambda ^{}\lambda }^{1/2}(\theta ^{}\theta ).$$
(B10)
With the aid of Eqs. (B5), (B8) and the identity
$$d_{\lambda ^{}\lambda }^{1/2}(\theta )d_{\lambda ^{}\lambda }^J(\theta )=\frac{1}{2}\left(P_{J{\scriptscriptstyle \frac{1}{2}}}(\mathrm{cos}\theta )+4\lambda \lambda ^{}P_{J+{\scriptscriptstyle \frac{1}{2}}}(\mathrm{cos}\theta )\right),$$
(B11)
we find the sought expression for the partial-wave amplitude $`T_r^{J\rho ^{}\rho }`$ with definite parity $`r`$ in terms of $`T_\pm `$:
$$T_r^{J\rho ^{}\rho }=\frac{1}{2}_1^1\mathrm{d}(\mathrm{cos}\theta )\left[T_r^{\rho ^{}\rho }(\theta )P_{J{\scriptscriptstyle \frac{1}{2}}}(\mathrm{cos}\theta )R_r^{\rho ^{}\rho }+T_r^{\rho ^{}\rho }(\theta )P_{J+{\scriptscriptstyle \frac{1}{2}}}(\mathrm{cos}\theta )R_r^{\rho ^{}\rho }\right],$$
(B12)
where the dependence on external momenta is omitted for brevity and only the dependence on the scattering angle is exhibited. The factors
$`R_r^{\rho ^{}\rho }=\{\begin{array}{cc}\rho \rho ^{}N_\rho N_\rho ^{}^{},\hfill & r=+\\ N_\rho N_\rho ^{}^{},\hfill & r=\end{array}`$ (B15)
$`N_\pm ={\displaystyle \frac{E_p\pm m_\mathrm{N}}{2E_p}},N_\pm ^{}={\displaystyle \frac{E_p^{}\pm m_\mathrm{N}}{2E_p^{}}},`$ (B16)
arise from the helicity spinors.
Now we can write down the relations due to the charge conjugation symmetry. It relates the amplitudes of positive energy with those of negative energy and opposite parity:
$$T_r^{J(\rho ^{},\rho )}(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0)=T_r^{J(\rho ^{},\rho )}(p_0^{},\mathrm{p}^{},p_0,\mathrm{p};P_0).$$
(B17)
Needless to say, these relations will only hold in the quasipotential formulations which satisfy the charge conjugation symmetry.
The isospin decomposition of the $`\pi N`$ amplitude is carried out as follows. If we denote $`\chi _N`$ and $`\varphi _\pi ^a`$ as the isospin states of the nucleon and pion, respectively, then
$`T`$ $`=`$ $`\varphi _{}^{}{}_{\pi }{}^{a}\chi _N^{}\left[\delta _{ab}T^{(+)}+i\epsilon _{abc}\tau _cT^{()}\right]\chi _N\varphi _\pi ^b`$ (B18)
$`=`$ $`\varphi _{}^{}{}_{\pi }{}^{a}\chi _N^{}\left[\frac{1}{3}\tau _a\tau _bT^{1/2}+(\delta _{ab}\frac{1}{3}\tau _a\tau _b)T^{3/2}\right]\chi _N\varphi _\pi ^b,`$ (B19)
where $`\tau _a`$ ($`a=1,2,3`$) are the isospin Pauli matrices, satisfying $`\frac{1}{2}[\tau _a,\tau _b]=i\epsilon _{abc}\tau _c.`$ Evidently,
$$T^{(+)}=\frac{1}{3}(T^{1/2}+2T^{3/2}),T^{()}=\frac{1}{3}(T^{1/2}T^{3/2}).$$
(B20)
## C The off-shell $`\pi N`$ potential
According to Eq. (1) the off-shell $`\pi N`$ amplitude can completely be specified by the scalar $`2\times 2`$ matrices $`A`$ and $`B`$. Here we give the expressions for these matrices corresponding to the tree-diagram potential in Fig. 2. Also the isospin factors, $`(I)`$, are given.
### 1 Baryon-exchange graphs
For the isospin-1/2 baryon: $`(\frac{1}{2})=1`$, $`(\frac{3}{2})=2`$.
For the isospin-3/2 baryon: $`(\frac{1}{2})=2`$, $`(\frac{3}{2})=\frac{1}{2}`$.
(a) The $`u`$-channel exchange of a baryon with spin $`\frac{1}{2}`$, mass $`m`$ and parity $`\eta `$, using the (pseudo-)scalar vs (pseudo-)vector admixture coupling, cf. Ref. , specified by parameter $`\lambda `$ \[$`\lambda =0`$ corresponds to pure (pseudo-)vector and is used in the text\]:
$`A^I`$ $`=`$ $`{\displaystyle \frac{g_{\pi NB}^2}{4\pi }}{\displaystyle \frac{1}{um^2}}[\lambda \left(\begin{array}{cc}m& \eta \\ \eta & 0\end{array}\right)+{\displaystyle \frac{\lambda (1\lambda )}{2m}}\left(\begin{array}{cc}2\eta u+\eta (p^2+p_{}^{}{}_{}{}^{2})& m\\ m& 2\eta \end{array}\right)`$ (C8)
$`+{\displaystyle \frac{(1\lambda )^2}{4m^2}}\left(\begin{array}{cc}m(up^2p_{}^{}{}_{}{}^{2})& \eta (p^2+p_{}^{}{}_{}{}^{2})\\ \eta (p^2+p_{}^{}{}_{}{}^{2})& m\end{array}\right)](I),`$
$`B^I`$ $`=`$ $`{\displaystyle \frac{g_{\pi NB}^2}{4\pi }}{\displaystyle \frac{1}{um^2}}[\lambda \left(\begin{array}{cc}\eta & 0\\ 0& 0\end{array}\right)+{\displaystyle \frac{\lambda (1\lambda )}{2m}}\left(\begin{array}{cc}2m& \eta \\ \eta & 0\end{array}\right)`$ (C16)
$`+{\displaystyle \frac{(1\lambda )^2}{4m^2}}\left(\begin{array}{cc}\eta u& m\\ m& 1\end{array}\right)](I).`$
(b) The $`\mathrm{\Delta }`$ exchange using the conventional coupling, Eq. (25):
$`A^I`$ $`=`$ $`{\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi }}{\displaystyle \frac{1}{um^2}}\{A_{P{\scriptscriptstyle \frac{3}{2}}}`$ (C17)
$`+`$ $`{\displaystyle \frac{2}{3m^2}}[a^2\left(\begin{array}{cc}0& p^{}p_u\overline{P}_{22}p_{}^{}{}_{}{}^{2}\\ pp_u\overline{P}_{22}p^2& 0\end{array}\right)`$ (C20)
$``$ $`2azm\left(\begin{array}{cc}\overline{P}_{22}(p_{}^{}{}_{}{}^{2}p+p^2p^{})p_u/u& 0\\ 0& 1(p+p^{})p_u/u\end{array}\right)`$ (C23)
$`+`$ $`m(\frac{1}{2}2az)\left(\begin{array}{cc}(p_{}^{}{}_{}{}^{2}k+p^2k^{})p_u/u(k+k^{})p_u2P_{22}& 0\\ 0& (k+k^{})p_u/u\end{array}\right)`$ (C26)
$`+`$ $`a(1+a)\left(\begin{array}{cc}0& k^{}p_u+2P_{22}\\ kp_u+2P_{22}& 0\end{array}\right)`$ (C29)
$``$ $`(1+a)^2P_{22}\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)m(1+2az)P_{22}\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)]\}(I),`$ (C34)
$`B^I`$ $`=`$ $`{\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi }}{\displaystyle \frac{1}{um^2}}\{B_{P{\scriptscriptstyle \frac{3}{2}}}`$ (C35)
$`+`$ $`{\displaystyle \frac{2}{3m^2}}[a^2\left(\begin{array}{cc}\overline{P}_{22}& 0\\ 0& 1\end{array}\right)2azm\left(\begin{array}{cc}0& p^{}p_u/u\\ pp_u/u& 0\end{array}\right)`$ (C40)
$`+`$ $`m(\frac{1}{2}2az)\left(\begin{array}{cc}0& kp_u/u\\ k^{}p_u/u& 0\end{array}\right)a(1+a)\left(\begin{array}{cc}(k^{}+k)p_u+2P_{22}& 0\\ 0& 0\end{array}\right)`$ (C45)
$`+`$ $`(1+a)^2P_{22}\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)]\}(I),`$ (C48)
where
$`a=z\frac{1}{2},`$ (C49)
$`P_{22}=(k^{}p_u)(kp_u)/u,\overline{P}_{22}=(p^{}p_u)(pp_u)/u,`$ (C50)
$`p_u=pk^{}=p^{}k,u=p_u^2,`$ (C51)
and $`A_{P{\scriptscriptstyle \frac{3}{2}}}`$, $`B_{P{\scriptscriptstyle \frac{3}{2}}}`$ are the contributions of the spin-3/2 projection operator.
For half-integer spin the contributions of the spin-$`j`$ projection operator read
$`A_{Pj}`$ $`=`$ $`{\displaystyle \frac{(1)^{j{\scriptscriptstyle \frac{1}{2}}}(j\frac{1}{2})!}{(2j)!!}}(\stackrel{~}{k}^2\stackrel{~}{k^{}}^2)^{(j{\scriptscriptstyle \frac{1}{2}})/2}[\left(\begin{array}{cc}m& 1\\ 1& 0\end{array}\right)P_{j+{\scriptscriptstyle \frac{1}{2}}}^{}(x_u)`$ (C57)
$`+\left(\begin{array}{cc}m[\overline{P}_{22}(p_{}^{}{}_{}{}^{2}p+p^2p^{})p_u/u]& p^{}p_u\overline{P}_{22}p_{}^{}{}_{}{}^{2}\\ pp_u\overline{P}_{22}p^2& 0\end{array}\right)P_{j{\scriptscriptstyle \frac{1}{2}}}^{}(x_u)],`$
$`B_{Pj}`$ $`=`$ $`{\displaystyle \frac{(1)^{j{\scriptscriptstyle \frac{1}{2}}}(j\frac{1}{2})!}{(2j)!!}}(\stackrel{~}{k}^2\stackrel{~}{k^{}}^2)^{(j{\scriptscriptstyle \frac{1}{2}})/2}[\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)P_{j+{\scriptscriptstyle \frac{1}{2}}}^{}(x_u)`$ (C63)
$`+\left(\begin{array}{cc}\overline{P}_{22}& mp^{}p_u/u\\ mpp_u/u& 1\end{array}\right)P_{j{\scriptscriptstyle \frac{1}{2}}}^{}(x_u)],`$
where $`P_l^{}(x)`$ is the first derivative of the Legendre polynomial, and
$$\stackrel{~}{k}=k\frac{(kp_u)}{u}p_u,\stackrel{~}{k}^{}=k^{}\frac{(k^{}p_u)}{u}p_u,x_u=\frac{\stackrel{~}{k}\stackrel{~}{k}^{}}{(\stackrel{~}{k}^2\stackrel{~}{k^{}}^2)^{{\scriptscriptstyle \frac{1}{2}}}}.$$
(C64)
(c) The $`\mathrm{\Delta }`$ exchange using the gauge-invariant coupling (26), according to Eq. (72) gives
$`A^I`$ $`=`$ $`{\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi m}}{\displaystyle \frac{u}{um^2}}A_{P{\scriptscriptstyle \frac{3}{2}}}(I),`$ (C65)
$`B^I`$ $`=`$ $`{\displaystyle \frac{f_{\pi N\mathrm{\Delta }}^2}{4\pi m_\pi m}}{\displaystyle \frac{u}{um^2}}B_{P{\scriptscriptstyle \frac{3}{2}}}(I).`$ (C66)
### 2 Meson-exchange graphs
For the isoscalar meson: $`(\frac{1}{2})=(\frac{3}{2})=2`$.
For the isovector meson: $`(\frac{1}{2})=2`$, $`(\frac{3}{2})=1`$.
(a) $`\rho `$-meson exchange:
$`A^I`$ $`=`$ $`{\displaystyle \frac{g_{\rho \pi \pi }g_{\rho NN}}{4\pi }}{\displaystyle \frac{1}{m_\rho ^2t}}\left(\begin{array}{cc}\frac{\kappa _\rho }{2m_N}(k^2+k_{}^{}{}_{}{}^{2}2s)& 1\\ 1& \frac{\kappa _\rho }{m_N}\end{array}\right)(I),`$ (C69)
$`B^I`$ $`=`$ $`{\displaystyle \frac{g_{\rho \pi \pi }g_{\rho NN}}{4\pi }}{\displaystyle \frac{1}{m_\rho ^2t}}\left(\begin{array}{cc}2& \frac{\kappa _\rho }{m_N}\\ \frac{\kappa _\rho }{m_N}& 0\end{array}\right)(I).`$ (C72)
(b) $`\sigma `$-meson exchange:
$`A^I`$ $`=`$ $`{\displaystyle \frac{g_{\sigma \pi \pi }g_{\sigma NN}}{8\pi m_\pi }}\left[c_\sigma {\displaystyle \frac{m_\pi ^2}{m_\sigma ^2}}{\displaystyle \frac{\frac{1}{2}(k^2+k_{}^{}{}_{}{}^{2})}{m_\sigma ^2t}}\right]\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)(I),`$ (C75)
$`B^I`$ $`=`$ $`0.`$ (C76)
## D $`\pi N`$ amplitudes for higher-spin baryon exchange
The use of GI couplings of higher-spin baryons allows us to treat exchanges of baryons with any spin in a straightforward way. Taking the point of view that consistent $`\pi NN_j^{}`$ couplings (where by $`N_j^{}`$ we denote the spin-$`j`$ baryon, $`j3/2`$) are those invariant under the appropriate gauge transformations of the $`N_j^{}`$ field, we can write down the following ansatz for the $`\pi N`$ scattering amplitude through a spin-$`j`$ baryon tree-level exchange:
$`M(k^{},k;P)`$ $`=`$ $`(1)^l{\displaystyle \frac{l!}{(2l+1)!!}}{\displaystyle \frac{g^2}{p/m}}`$ (D1)
$`\times `$ $`\left[P_{l+1}^{}(\widehat{k}\widehat{k}^{})+\widehat{k}/^{}\widehat{k}/P_l^{}(\widehat{k}\widehat{k}^{})\right]\left(P^2\sqrt{\stackrel{~}{k}^2\stackrel{~}{k^{}}^2}\right)^l,`$ (D2)
where $`l=j\frac{1}{2}`$, $`P_l^{}(x)`$ is the first derivative of the Legendre polynomial, $`P`$ is the momentum of the exchanged baryon, and
$$\stackrel{~}{k}_\mu =k_\mu \frac{kP}{P^2}P_\mu ,\widehat{k}_\mu =\stackrel{~}{k}_\mu /\sqrt{\stackrel{~}{k}^2}.$$
(D3)
This amplitude is actually just the spin-$`j`$ projection operator contracted with the external pion momenta and multiplied by $`g^2p^{2l}(p/m)^1`$. Because of the projection operator, the amplitude contains only the spin-$`j`$ contributions, as we will now demonstrate.
In the CMS $`P=(W,\stackrel{}{0})`$, hence $`\stackrel{~}{k}=(0,\stackrel{}{k})=(0,\stackrel{}{p})`$. The $`s`$-channel helicity amplitude is then written as,
$`M_{\lambda ^{}\lambda }^{\rho ^{}\rho }`$ $`=`$ $`\overline{u}_\lambda ^{}^\rho ^{}(\stackrel{}{p}^{})M(k^{},k;p)u_\lambda ^\rho (\stackrel{}{p})=C_j\overline{u}_\lambda ^{}^\rho ^{}(p^{}){\displaystyle \frac{W\gamma _0+m}{sm^2}}[P_{l+1}^{}(x_s)`$ (D4)
$`+`$ $`(\mathrm{pp}^{})^1(\stackrel{}{\gamma }\stackrel{}{p}^{})(\stackrel{}{\gamma }\stackrel{}{p})P_l^{}(x_s)](\mathrm{pp}^{}s)^lu^\rho _\lambda (\stackrel{}{p}),`$ (D5)
where $`x_s=\mathrm{cos}\theta ,`$ $`\theta `$ is the c.m. scattering angle, $`s=W^2`$, and the constant factors are absorbed in $`C_j`$. Using the Dirac equation, $`(\stackrel{}{\gamma }\stackrel{}{p})u_\lambda ^\rho (p)=(\rho E_p\gamma _0m_N)u_\lambda ^\rho (p),`$ we obtain
$`M_{\lambda ^{}\lambda }^{\rho ^{}\rho }`$ $`=`$ $`C_j\overline{u}_\lambda ^{}^\rho ^{}(p^{})[{\displaystyle \frac{W\gamma _0+m}{sm^2}}P_{l+1}^{}(x_s)`$ (D6)
$`+`$ $`{\displaystyle \frac{W\gamma _0+m}{sm^2}}{\displaystyle \frac{\rho ^{}E_p^{}\gamma _0m_N}{\mathrm{p}^{}}}{\displaystyle \frac{\rho E_p\gamma _0m_N}{\mathrm{p}}}P_l^{}(x_s)](\mathrm{pp}^{}s)^lu^\rho _\lambda (p).`$ (D7)
The contribution of this amplitude to the partial wave with total spin $`J`$ is given by the corresponding partial-wave amplitude. The latter can be obtained by the procedure of Appendix B, for that we only need to know the following angular integrals:
$$\frac{1}{2}_1^1dxP_l^{}(x)P_L(x)=\{\begin{array}{cc}1,\hfill & L=l2n1\hfill \\ 0,\hfill & \mathrm{otherwise}\hfill \end{array},$$
(D8)
where $`n`$ should be an integer between 0 and $`(l1)/2`$. The resulting parity-conserving partial-wave amplitude is then found to be
$`M_r^{\rho ^{}\rho J}`$ $`=`$ $`C_j(\mathrm{pp}^{}s)^l\times \{\begin{array}{cc}S^{(+r)}R_r^{\rho ^{}\rho },\hfill & J=j,\hfill \\ S^{(r)}(R_r^{\rho ^{}\rho }+๐ฉ_r^{}๐ฉ_rR_r^{\rho ^{}\rho }),\hfill & J=j12n,\hfill \\ S^{(+r)}(R_r^{\rho ^{}\rho }+๐ฉ_r^{}๐ฉ_rR_r^{\rho ^{}\rho }),\hfill & J=j22n,\hfill \end{array}`$ (D12)
where $`๐ฉ_\pm =(\rho E_p\pm m_N)/\mathrm{p}`$, $`S^{(\pm )}=(\pm Wm)^1`$, and factors $`R`$ are defined in Eq. (B15). Using the explicit form of $`R`$, we can see that the lower partial-wave contributions vanish exactly. Thus, amplitude Eq. (D1) has only the highest-spin contribution:
$$M_r^{\rho ^{}\rho J}=\delta _{jJ}C_j(\mathrm{pp}^{}s)^lS^{(r)}R_r^{\rho ^{}\rho }.$$
(D13) |
warning/0003/cond-mat0003397.html | ar5iv | text | # Generalized model for dynamic percolation
## I Introduction.
The percolation concept has turned out very useful for understanding transport and conduction processes in a wide range of disordered media, as exemplified by ionic conduction in polymeric, amorphous or glassy ceramic electrolytes, diffusion in biological tissues and permeability of disordered membranes .
Most of the situations discussed in Refs. pertain, however, to systems with โfrozenโ disorder; that is, the random environment in which a given transport process takes place does not change in time. This is certainly the case in many instances, but it is not true in general. As a matter of fact, there are many experimental systems in which the static percolation picture does not apply since the structure of the host material undergoes essential structural reorganizations on a time scale comparable to that at which the transport itself occurs. A few stray examples of such systems include certain biomembranes , solid protonic conductors , oil-continuous microemulsions and polymer electrolytes .
More specifically, ionic transport across a biomembrane, such as, e.g. gramicidin-$`A`$, occurs by the motion of ions through molecular channels along which they encounter potential barriers that fluctuate in time. The fluctuations of potential barriers may hinder significantly the transport and constitute an important transport-controlling factor . In the case of protonic conduction by the Grotthus mechanism , site-to-site motion of carriers occurs only between those neighboring $`H_2O`$ or $`NH_3`$ groups that have a favorable relative orientation; thermally activated rotation of these groups is the structural host-reorganization process interacting with the carrier motion. Similarly, within oil-continuous microemulsions, the charge transport proceeds by charge being transfered from one water globule to another, as globules approach each other in their Brownian motion . Lastly, in polymer electrolytes, such as, e.g. polyethylene oxide complexed nonstoichiometrically with the ionic salt $`NaSCN`$, the $`Na^+`$ ions are largely tetrahedrally coordinated by polyether oxygenes, but at the same time that $`Na^+`$ ions hop from one fourfold coordination site to another, the oxygens themselves, along with the polymeric backbone, undergo large-amplitude wagging and even diffusive motion .
Clearly, all the above mentioned examples involve two characteristic time scales, one which describes the typical time $`\tau `$ between two successive hops of the carrier, while the other is associated with a typical renewal time $`\tau ^{}`$ of the environment itself; namely, the time needed for the host medium to re-organize itself and thereby provide a new set of available pathways for transport. Consequently, the static percolation picture applies only when the characteristic time $`\tau ^{}`$ gets infinitely large. For a finite $`\tau ^{}`$ $`dynamicpercolation`$ has to be considered, and one encounters quite a different behavior when compared to the random environments with quenched disorder. As a result, one observes Ohmic-type or Stokes-type linear velocity-force relation for the carrierโs terminal velocities as a function of the applied field, in contrast to the threshold behavior predicted by the static percolation theory. The prefactor in the linear velocity-force relation may depend, however, in a non-trivial way on the systemโs parameters and this dependence consitutes the main challenge for the theoretical analysis here. On the other hand, we note that in the above mentioned examples of the dynamic percolative environments quite different physical processes are responsible for the time evolution of the host medium. Consequently, one expects that the prefactor in the Stokes-type velocity-force relation should also be dependent on the precise mechanism which underlies the temporal re-organization of the environment.
Theoretical modelling of dynamic percolative environments has followed several avenues, which differ mostly in how the time evolution of the disorder is constrained; Namely, is it constrained (a) by conservation laws or (b) by spatial and temporal correlations in the renewal events? Early models of dynamic percolation described the random environment within the framework of a standard bond-percolation model, in which the strength of each bond fluctuates in time between zero and a finite value. The dynamics of the host medium in these models was accounted for by a series of instantaneous renewal events. These events were assumed to occur at random times, chosen from a renewal time distribution. In the renewal process the positions of all unblocked bonds are being reassigned, such that after each renewal event a carrier sees a newly defined network. This approach is thus characterized by a $`globaldynamicaldisorder`$ without global conservation laws and correlations, since the entire set of random hopping rates is simultaneously renewed independently of the previous history. Another model characterized by a $`localdynamicaldisorder`$ has been proposed in Refs. and , and subsequently generalized to the non-Markovian case in Ref.. This model appears to be similar to the previous one, except that here the hopping rates at different sites fluctuate $`independently`$ of each other. That is, individual bonds, rather than the whole lattice change in the renewal events. To describe the dynamical behavior in the $`localdynamicaldisorder`$ case, a dynamical mean-field theory has been proposed , based on the effective medium approximation introduced for the analysis of random walks on lattices with static disorder , and has been generalized to include the possibility of multistate transformations of the $`dynamically`$ random medium . More recently, several exactly solvable one-dimensional models with $`global`$ and $`local`$ dynamical disorder have been discussed .
In the second approach, which emerged within the context of the ionic conductivity in superionic solids, the dynamical percolative environment has been considered as a multicomponent mixture of mobile species in which one or several neutral components block the carrier component . In particular, such a situation can be observed in a superionic conductor $`\beta ^{\prime \prime }`$-alumina, doped with two different ionic species (e.g. $`Na^+`$ and $`Ba^{2+}`$), where small $`Na^+`$ ions are rather mobile, while the larger $`Ba^{2+}`$ ions move essentially slower and temporarily block the $`Na^+`$ ions. Contrary to the previous line of thought, the dynamics of such a percolative environment has essential correlations, generated by hard-core exclusion interactions between the species involved, and moreover, it obeys the conservation law \- the total number of the particles involved is conserved. In Ref., the frequency-dependent ionic conductivity of the light species has been analysed combining a continuous time random walk approach for the dynamical problem with an effective medium approximation describing the frozen environment of slow species. Next, as an explanation of the sharp increase of electrical conductivity transition in water-in-oil microemulsions when the volume fraction of water is increased towards a certain threshold value, in Refs. and it has been proposed that the charge carriers are not trapped in the finite water clusters, but rather a charge on a water globule can propagate by either hopping to a neighboring globule, when they approach each other, or via the diffusion of the host globule itself. This picture has been interpreted in terms of a model similar to that employed in Ref., with the only difference being that here the โblockersโ of Ref. play the role of the transient charge carriers. In the model of Refs. and , in which the host dynamics is influenced by spatial correlations and conservation of the number of the water globules involved, the conductivity depends hence, on the rate of cluster rearrangement. Lastly, a similar problem of a carrier diffusion in an environment created by mobile hard-core lattice-gas particles has been analysed in Ref. by using the developed dynamic bond percolation theory of Refs. and .
In this paper we propose a generalized model of dynamic percolation which shares common features with both bond-fluctuating models of Refs. as well as models involving mobile blockers of Refs.. The system we consider consists of a host lattice, which here is a regular cubic lattice whose sites support at most a single occupancy, hard-core โenvironmentโ particles, and a single hard-core carrier particle. The โenvironmentโ particles move on the lattice by performing a random hopping between the neighboring lattice sites, which is constrained by the hard-core interactions, and may disappear from and re-appear (renewal processes) on the empty sites of the lattice with some prescribed rates<sup>*</sup><sup>*</sup>*We hasten to remark that diffusive processes, of course, also result in a certain renewal of the environment; diffusive processes, as compared to the spontaneous creation and annihilation of particles, have however completelely different underlying physics and influence in a completely different fashion the evolution of the system, as we proceed to show. Following the terminology of Refs., we thus choose here to distinguish between diffusive and creation/annihilation processes, referring to the latter as the renewal ones.. In turn, the carrier particle is always present on the lattice, i.e., it can not disappear spontaneously, and is subject to a constant external force $`\stackrel{}{E}`$. Hence, the carrier performs a biased random walk, which is constrained by the hard-core interactions with the โenvironmentโ particles, and probes the response of the percolative environment to the internal perturbancy or, in other words, the frictional properties of such a dynamical environment.
An important aspect of our model, which makes it different to the previously proposed models of dynamic percolation, is that we include the hard-core interaction between โenvironmentโ particles and the carrier molecule, such that the latter may influence the dynamics of the environment. This results, as we proceed to show, in the emergence of complicated density profiles of the โenvironmentโ particles around the carrier. These profiles, as well as the terminal velocity $`V_c`$ of the carrier, are determined here explicitly, in terms of an approximate approach of Ref., which is based on the decoupling of the carrier-particle-particle correlation functions into the product of pair-wise correlations. We show that the โenvironmentโ particles tend to accumulate in front of the driven carrier creating a sort of a โtraffic jamโ, which impedes its motion. Thus the density profiles around the carrier are highly asymmetric: the local density of the โenvironmentโ particles in front of the carrier is higher than the average and approaches the average value as an exponential function of the distance from the carrier. The characteristic length and the amplitude of the density relaxation function are calculated explicitly. On the other hand, past the carrier the local density is lower than the average: We show that depending on the condition whether the number of particles in the percolative environment is explicitly conserved or not, the local density past the carrier may tend to the average value either as an exponential or even as an $`\mathrm{a}\mathrm{l}\mathrm{g}\mathrm{e}\mathrm{b}\mathrm{r}\mathrm{a}\mathrm{i}\mathrm{c}`$ function of the distance, revealing in the latter case especially strong memory effects and strong correlations between the particle distribution in the environment and the carrier position. Further on, we find that the terminal velocity of the carrier particle depends explicitly on the excess density in the โjammedโ region in front of the carrier, as well as on the โenvironmentโ particles density past the carrier. Both, in turn, are dependent on the magnitude of the velocity, as well as on the rate of the renewal processes and the rate at which the โenvironmentโ particles can diffuse away from the carrier. The interplay between the jamming effect of the environment, produced by the carrier particle, and the rate of its homogenization due to diffusive smoothening and renewal processes, manifests itself as a medium-induced frictional force exerted on the carrier, whose magnitude depends on the carrier velocity. As a consequence of such a non-linear coupling, in the general case, (i.e. for arbitrary rates of the renewal and diffusive processes), $`V_c`$ can be found only implicitly, as the solution of a non-linear equation relating $`V_c`$ to the system parameters. This equation simplifies considerably in the limit of small applied external fields $`\stackrel{}{E}`$ and we find that the force-velocity relation to the field becomes linear. This implies that the frictional force exerted on the carrier particle by the environment is $`\mathrm{v}\mathrm{i}\mathrm{s}\mathrm{c}\mathrm{o}\mathrm{u}\mathrm{s}`$. This linear force-velocity relation can be therefore interpreted as the analog of the Stokes formula for the dynamic percolative environment under study; in this case, the carrier velocity is calculated explicitly as well as the corresponding friction coefficient. In turn, this enable us to estimate the self-diffusion coefficient of the carrier in absence of external field; we show that when only diffusive re-arrangement of the percolative environment is allowed, while the renewal processes are suppressed, the general expression for the diffusion coefficient reduces to the one obtained previously in Refs. and . We note that the result of Refs. and is known to serve as a very good approximation for the self-diffusion coefficient in hard-core lattice-gases .
We finally remark, that a qualitatively similar physical effect was predicted recently for a different model system involving a charged particle moving at a constant speed a small distance above the surface of an incompressible, infinitely deep liquid. It has been shown in Refs., that the interactions between the moving particle and the fluid molecules induce an effective frictional force exerted on the particle, producing a local distortion of the liquid interface, - a bump, which travels together with the particle and increases effectively its mass. The mass of the bump, which is analogous to the jammed region appearing in our model, depends itself on the particleโs velocity resulting in a non-linear coupling between the medium-induced frictional force exerted on the particle and its velocity .
The paper is structured as follows: In Section II we formulate the model and introduce basic notations. In Section III we write down the dynamical equations which govern the time evolution of the โenvironmentโ particles and of the carrier. Section IV is devoted to the analytical solution of these evolution equations in the limit $`t\mathrm{}`$ ; here we also present some general results on the shape of the density profiles around stationary moving carrier and on the carrier terminal velocity, which is given implicitly, as the solution of a transcendental equation defining the general force-velocity relation for the dynamic percolative environment under study. In Section V we derive explicit asymptotic results for the carrier terminal velocity in the limit of small applied external fields $`\stackrel{}{E}`$ and obtain the analog of the Stokes formula for such a percolative environment; as well, we present here explicit results for the friction coefficient of the host medium and for the self-diffusion coefficient of the carrier in the absence of external field. Asymptotic behavior of the density profiles of the โenvironmentโ particles around the carrier is discussed in Section VI. Finally, we conclude in Section VII with a brief summary and discussion of our results.
## II The model.
The model for dynamic percolation we study here consists of a three-dimensional simple cubic lattice of spacing $`\sigma `$, the sites of which are partially occupied by identical hard-core โenvironmentโ particles and a single, hard-core, carrier particle (see Fig.1). For both types of particles the hard-core interactions prevent multiple occupancy of the lattice sites; that is, no two โenvironmentโ particles or the โcarrierโ and an โenvironmentโ particle can occupy simultaneously the same site, and particles can not pass through each other.
The occupation of the lattice sites by the โenvironmentโ particles is characterized by the time-dependent occupation variable $`\eta (\stackrel{}{r})`$, $`\stackrel{}{r}`$ being the lattice-vector of the site in question. This variable assumes two values:
$$\eta (\stackrel{}{r})=\{\begin{array}{cc}1,\text{ if the site }\stackrel{}{r}\text{ is occupied}\hfill & \\ 0,\text{ if the site }\stackrel{}{r}\text{ is empty}\hfill & \end{array}$$
(1)
Next, we assume the following dynamics of the โenvironmentโ particles: The particles can spontaneously disappear from the lattice, and may re-appear at random positions and random time moments, which is reminiscent of the host medium dynamics stipulated in Refs.. We refer to these two processes generally as renewal processes. In addition, the environment particles move randomly within the lattice by performing nearest-neighbor random walks constrained by the hard-core interactions, which is the main feature of the approach in Refs.. We stipulate that any of the โenvironmentโ particles waits a time $`\delta \tau `$, which has an exponential probability distribution with a mean $`\tau ^{}`$, and then chooses from a few possibilities: (a) disappearing from the lattice at rate $`g`$, which is realized instantaneously, or (b) attempting to hop, at rate $`l/6`$, onto one of $`6`$ neighboring sites. The hop is actually fulfilled if the target site is not occupied at this time moment by any other particle; otherwise, the particle attempting to hop remains at its initial position, and (c) particles may re-appear on any $`\mathrm{v}\mathrm{a}\mathrm{c}\mathrm{a}\mathrm{n}\mathrm{t}`$ lattice site with rate $`f`$.
Note that, for simplicity, we assumed that the characteristic diffusion time and the renewal times of the โenvironmentโ particles are equal to each other. These times, i.e. $`\tau _{dif}`$, mean creation time $`\tau _{cr}`$ and mean annihilation time $`\tau _{an}`$ may, however, be different, and can be restored in our final results by a mere replacement $`ll\tau ^{}/\tau _{dif}`$, $`ff\tau ^{}/\tau _{cr}`$ and $`gg\tau ^{}/\tau _{an}`$.
Note also that the number of particles is not explicitly conserved in such a dynamical model of the environment, which happens because of the presence of the renewal processes; the particles diffusion, on contrary, conserves the particles number. However, in the absence of attractive particle-particle interactions and external perturbances, the particles distribution on the lattice is uniform and the average occupation $`\rho (t)=\overline{\eta (\stackrel{}{r})}`$ of the lattice tends, as $`t\mathrm{}`$, to a constant value, $`\rho _s=f/(f+g)`$. This relation can be thought of as the Langmuir adsorption isotherm .
Hence, the limit $`\tau _{dif}\mathrm{}`$ (or, $`l0`$) corresponds to the ordinary site percolation model with immobile blocked sites. The limit $`f,g0`$, ($`\tau _{cr},\tau _{an}\mathrm{}`$), while keeping the ratio $`f/g`$ fixed, $`f/g=\rho _s/(1\rho _s)`$, corresponds to the usual hard-core lattice-gas with the number of particles conserved.
At time $`t=0`$ we introduce at the origin of the lattice an extra particle, the carrier, whose role is to probe the response of the environment modeled by dynamic percolation to an external perturbance. We stipulate that only the carrier out of all participating particles can not disappear from the system, and moreover, its motion is biased by some external constant force. As a physical realization, we envisage that the carrier is charged, while all other particles are neutral, and the system is exposed to constant external electric field $`\stackrel{}{E}`$. The dynamics of the carrier particle is defined as follows: We suppose that the waiting time between successive jumps of the carrier has also an exponential distribution with a mean value $`\tau `$, which may in general be different from the corresponding waiting time of the environment particles. Attempting to hop, the carrier first chooses a hop direction with probabilities
$$p_\mu =exp\left[\frac{\beta }{2}\left(\stackrel{}{E}\stackrel{}{e}_\mu \right)\right]/\underset{\nu }{}exp\left[\frac{\beta }{2}\left(\stackrel{}{E}\stackrel{}{e}_\nu \right)\right],$$
(2)
where $`\beta `$ is the reciprocal temperature, $`\stackrel{}{e}_\nu `$ (or $`\stackrel{}{e}_\mu `$) stand for six unit lattice vectors, $`\nu ,\mu =\{\pm 1,\pm 2,\pm 3\}`$, connecting the carrier position with $`6`$ neighboring lattice sites, and $`(\stackrel{}{E}\stackrel{}{e}_\nu )`$ denotes the scalar product. We adopt the convention that $`\pm 1`$ corresponds to $`\pm X`$, $`\pm 2`$ corresponds to $`\pm Y`$ while $`\pm 3`$ stands for $`\pm Z`$. The jump is actually fulfilled when the target lattice site is vacant. Otherwise, as mentioned the carrier remains at its position. For simplicity we assume in what follows that the external field is oriented along the $`X`$-axis in the positive direction, such that $`\stackrel{}{E}=(E,0,0)`$. Note also that for the choice of the transition probabilities as in Eq.(2), the detailed balance is naturally preserved.
## III Evolution Equations
Let $`P(\stackrel{}{R}_c,\eta ;t)`$ denote the joint probability that at time moment $`t`$ the carrier occupies position $`\stackrel{}{R}_c`$ and all โenvironmentโ particles are in configuration $`\eta \{\eta (\stackrel{}{r})\}`$. Next, let $`\eta ^{\stackrel{}{r},\mu }`$ denote particlesโ configuration obtained from $`\eta `$ by exchanging the occupation variables of the sites $`\stackrel{}{r}`$ and $`\stackrel{}{r}+\stackrel{}{e}_\stackrel{}{\mu }`$, i.e. $`\eta (\stackrel{}{r})\eta (\stackrel{}{r}+\stackrel{}{e}_\stackrel{}{\mu })`$, and $`\widehat{\eta }^\stackrel{}{r}`$ be the configuration obtained from $`\eta `$ by changing the occupation of the site $`\stackrel{}{r}`$ as $`\eta (\stackrel{}{r})1\eta (\stackrel{}{r})`$. Clearly, the first type of process appears due to random hops of the โenvironmentโ particles, while the second one stems from the renewal processes, i.e. random creation and annihilation of the โenvironmentโ particles. Then, summing up all possible events which can result in the configuration $`(\stackrel{}{R}_c,\eta )`$ or change this configuration for any other, we find that the temporal evolution of the system under study is governed by the following master equation:
$`_tP(\stackrel{}{R}_c,\eta ;t)={\displaystyle \frac{l}{6\tau ^{}}}{\displaystyle \underset{\mu }{}}{\displaystyle \underset{\stackrel{}{r}\stackrel{}{R}_c\stackrel{}{e}_\stackrel{}{\mu },\stackrel{}{R}_c}{}}\left\{P(\stackrel{}{R}_c,\eta ^{\stackrel{}{r},\mu };t)P(\stackrel{}{R}_c,\eta ;t)\right\}+`$ (3)
$`+`$ $`{\displaystyle \frac{1}{\tau }}{\displaystyle \underset{\mu }{}}p_\mu \left\{\left(1\eta (\stackrel{}{R}_c)\right)P(\stackrel{}{R}_c\stackrel{}{e}_\stackrel{}{\mu },\eta ;t)\left(1\eta (\stackrel{}{R}_c+\stackrel{}{e}_\stackrel{}{\mu })\right)P(\stackrel{}{R}_c,\eta ;t)\right\}+`$ (4)
$`+`$ $`{\displaystyle \frac{g}{\tau ^{}}}{\displaystyle \underset{\stackrel{}{r}\stackrel{}{R}_c}{}}\left\{\left(1\eta (\stackrel{}{r})\right)P(\stackrel{}{R}_c,\widehat{\eta }^\stackrel{}{r};t)\eta (\stackrel{}{r})P(\stackrel{}{R}_c,\eta ;t)\right\}+`$ (5)
$`+`$ $`{\displaystyle \frac{f}{\tau ^{}}}{\displaystyle \underset{\stackrel{}{r}\stackrel{}{R}_c}{}}\left\{\eta (\stackrel{}{r})P(\stackrel{}{R}_c,\widehat{\eta }^\stackrel{}{r};t)\left(1\eta (\stackrel{}{r})\right)P(\stackrel{}{R}_c,\eta ;t)\right\}.`$ (6)
Note that the terms in the first (resp. second) line of Eq.(6) describe random hopping motion of the โenvironmentโ particles (resp. biased motion of the carrier) in terms of the Kawasaki-type particle-vacancy exchanges, while the terms in the third and the fourth lines account for the Glauber-type decay and creation of the โenvironmentโ particles.
### A Mean velocity of the carrier and correlation functions.
From Eq.(6) we can readily compute the velocity of the carrier. Multiplying both sides of Eq.(6) by $`(\stackrel{}{R}_c\stackrel{}{e}_1)`$ and summing over all possible configurations $`(\stackrel{}{R}_c,\eta )`$ we find that the carrierโs mean velocity $`V_c(t)`$, defined as
$$V_c(t)\frac{\mathrm{d}}{\mathrm{dt}}\left(\overline{\stackrel{}{R}_c\stackrel{}{e}_1}\right),$$
(7)
obeys:
$$V_c(t)=\frac{\sigma }{\tau }\left\{p_1(1k(\stackrel{}{e}_1;t))p_1(1k(\stackrel{}{e}_1;t))\right\},$$
(8)
where $`k(\stackrel{}{\lambda };t)`$ stands for the carrier-โenvironmentโ particles pair correlation function
$$k(\stackrel{}{\lambda };t)\underset{\stackrel{}{R}_c,\eta }{}\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })P(\stackrel{}{R}_c,\eta ;t).$$
(9)
In other words, $`k(\stackrel{}{\lambda };t)`$ can be thought of as the density distribution of the โenvironmentโ particles as seen from the carrier which moves with velocity $`V_c(t)`$.
Hence, $`V_c(t)`$ depends explicitly on the local density of the โenvironmentโ particles in the immediate vicinity of the carrier. Note that if the โenvironmentโ is perfectly homogeneous, i.e. if for any $`\stackrel{}{\lambda }`$ the density profile is constant, $`k(\stackrel{}{\lambda };t)=\rho _s`$, which immediately implies decoupling between $`\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })`$ and $`P(\stackrel{}{R}_c,\eta ;t)`$ in Eq.(9), then we obtain from Eq.(8) a trivial mean-field-type result
$$V_c^{(0)}=(p_1p_1)(1\rho _s)\frac{\sigma }{\tau },$$
(10)
which states that the frequency of jumps of the carrier particles ($`\tau ^1`$) only gets renormalized by a factor $`1\rho _s`$, which gives the fraction of successful jumps.
The salient feature of our model is that there are essential backflow effects. The carrier effectively perturbs the spatial distribution of the โenvironmentโ particles so that stationary density profiles emerge. This can be contrasted to the earlier dynamic percolation models in which the carrier had no impact on the embedding medium and hence there was no re-arrengement of the host medium around the carrier particle. As a consequence $`k(\stackrel{}{\lambda };t)\rho _s`$, and $`k(\stackrel{}{\lambda };t)`$ approaches $`\rho _s`$ only at infinite separations from the carrier, i.e. when $`|\stackrel{}{\lambda }|\mathrm{}`$. Therefore, we rewrite Eq.(8) in the form
$$V_c(t)=V_c^{(0)}\frac{\sigma }{\tau }\{p_1(k(\stackrel{}{e}_1;t)\rho _s)p_1(\rho _sk(\stackrel{}{e}_1;t))\},$$
(11)
which shows explicitly the deviation of the mean velocity of the carrier from the mean-field-type result in Eq.(10) due to the formation of the density profiles.
### B Evolution equations of the pair correlation functions.
From Eq.(8) it follows that in order to obtain $`V_c(t)`$, it suffices to compute $`k(\stackrel{}{e}_{\pm 1};t)`$. Consequently, we have to evaluate the equation governing the time evolution of the pair correlation functions. Multiplying both sides of Eq.(6) by $`\eta (\stackrel{}{R}_c)`$ and summing over all configurations $`(\stackrel{}{R}_c,\eta )`$, we find that $`k(\stackrel{}{\lambda };t)`$ obeys
$`_tk(\stackrel{}{\lambda };t)`$ $`=`$ $`{\displaystyle \frac{l}{6\tau ^{}}}{\displaystyle \underset{\mu }{}}(_\mu \delta _{\stackrel{}{\lambda },\stackrel{}{e}_\mu }_\mu )k(\stackrel{}{\lambda };t){\displaystyle \frac{(f+g)}{\tau ^{}}}k(\stackrel{}{\lambda };t)+{\displaystyle \frac{f}{\tau ^{}}}+`$ (12)
$`+`$ $`{\displaystyle \frac{1}{\tau }}{\displaystyle \underset{\mu }{}}{\displaystyle \underset{\stackrel{}{R}_c,\eta }{}}p_\mu \left(1\eta (\stackrel{}{R}_c+e_\mu )\right)_\mu \eta (\stackrel{}{R}_c+\stackrel{}{\lambda })P(\stackrel{}{R}_c,\eta ;t),`$ (13)
where $`_\mu `$ denotes the ascending finite difference operator of the form
$$_\mu f(\stackrel{}{\lambda })f(\stackrel{}{\lambda }+\stackrel{}{e}_\mu )f(\stackrel{}{\lambda }),$$
(14)
and
$$\delta _{\stackrel{}{r},\stackrel{}{r}^{}}=\{\begin{array}{cc}1,\text{ if the site }\stackrel{}{r}=\stackrel{}{r}^{}\hfill & \\ 0,\text{ otherwise.}\hfill & \end{array}$$
(15)
The Kroneker-delta term $`\delta _{\stackrel{}{\lambda },\stackrel{}{e}_\mu }`$ signifies that the evolution of the pair correlations, Eq.(13), proceeds differently at large separations and at the immediate vicinity of the carrier. This stems from the asymmetric hopping rules of the carrier particle defined by Eq.(2).
Note next that the contribution in the second line in Eq.(13), which is associated with the biased diffusion of the carrier, appears to be non-linear with respect to the occupation numbers, such that the pair correlation function gets effectively coupled to the evolution of the third-order correlations of the form
$$T(\stackrel{}{\lambda },\stackrel{}{e}_\nu ;t)\underset{\stackrel{}{R}_c,\eta }{}\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })\eta (\stackrel{}{R}_c+\stackrel{}{e}_\nu )P(\stackrel{}{R}_c,\eta ;t).$$
(16)
That is, Eq.(13) is not closed with respect to the pair correlations but rather represents a first equation in the infinite hierachy of coupled equations for higher-order correlation functions. One faces, therefore, the problem of solving an infinite hierarchy of coupled differential equations and needs to resort to an approximate closure scheme.
### C Decoupling Approximation
Here we resort to the simplest non-trivial closure approximation, based on the decoupling of the third-order correlation functions into the product of pair correlations. More precisely, we assume that for $`\stackrel{}{\lambda }\stackrel{}{e}_\nu `$, the third-order correlation fulfils
$`{\displaystyle \underset{\stackrel{}{R}_c,\eta }{}}\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })\eta (\stackrel{}{R}_c+\stackrel{}{e}_\nu )P(\stackrel{}{R}_c,\eta ;t)`$ (17)
$``$ $`\left({\displaystyle \underset{\stackrel{}{R}_c,\eta }{}}\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })P(\stackrel{}{R}_c,\eta ;t)\right)\left({\displaystyle \underset{\stackrel{}{R}_c,\eta }{}}\eta (\stackrel{}{R}_c+\stackrel{}{e}_\nu )P(\stackrel{}{R}_c,\eta ;t)\right),`$ (18)
or, in other words,
$$\underset{\stackrel{}{R}_c,\eta }{}\eta (\stackrel{}{R}_c+\stackrel{}{\lambda })\eta (\stackrel{}{R}_c+\stackrel{}{e}_\nu )P(\stackrel{}{R}_c,\eta ;t)k(\stackrel{}{\lambda };t)k(\stackrel{}{e}_\nu ;t)$$
(19)
The approximate closure in Eq.(19) has been already employed for studying related models of biased carrier diffusion in hard-core lattice gases and has been shown to provide quite an accurate description of both the dynamical and stationary-state behavior. The decoupling in Eq.(19) was first introduced in Ref. to determine the properties of a driven carrier diffusion in a one-dimensional hard-core lattice gas with a conserved number of particles, i.e. without an exchange of particles with the reservoir. Extensive numerical simulations performed in Ref. have demonstrated that such a decoupling is quite a plausible approximation for the model under study. Moreover, rigorous probabilistic analysis of Ref. has shown that for this model the results based on the decoupling scheme in Eq.(19) are exact. Furthermore, the same closure procedure has been recently applied to study spreading of a hard-core lattice gas from a reservoir attached to one of the lattice sites . Again, a very good agreement between the analytical results and the numerical data has been found. Next, the decoupling in Eq.(19) has been used in a recent analysis of a biased carrier dynamics in a one-dimensional model of an adsorbed monolayer in contact with a vapour phase , i.e. a one-dimensional version of the model to be studied here. Also in this case an excellent agreement has been observed between the analytical predictions and the Monte Carlo simulations data . We now show that the approximate closure of the hierarchy of the evolution equations in Eq.(19) allows us to reproduce in the limit $`f,g=0`$ and $`f/g=const`$ the results of Refs. and , which are known (see e.g. Ref.) to provide a very good approximation for the carrier diffusion coefficient in three-dimensional hard-core lattice gases with arbitrary particle density. We expect therefore that such a closure scheme will render a plausible description of the carrier dynamics in a three-dimensional generalized dynamic percolation model. We base our further analysis on this approximation.
Making use of Eq.(19), we find from Eq.(13) that the pair correlations obey the following equations:
$$_tk(\stackrel{}{\lambda };t)=\frac{l}{6\tau ^{}}\stackrel{~}{L}k(\stackrel{}{\lambda };t)+\frac{f}{\tau ^{}},$$
(20)
which hold for all lattice sites except for those at the immediate vicinity of the carrier, i.e. for all $`\stackrel{}{\lambda }`$ except for $`\stackrel{}{\lambda }=\{\mathrm{๐},\stackrel{}{e}_{\pm 1},\stackrel{}{e}_{\pm 2},\stackrel{}{e}_{\pm 3}\}`$, while at the sites adjacent to the carrier one has
$$_tk(\stackrel{}{e}_\nu ;t)=\frac{l}{6\tau ^{}}\left(\stackrel{~}{L}+A_\nu (t)\right)k(\stackrel{}{e}_\nu ;t)+\frac{f}{\tau ^{}},$$
(21)
where $`\nu =\{\pm 1,\pm 2,\pm 3\}`$. The operators $`\stackrel{~}{L}`$ and coefficients $`A_\nu (t)`$ are given explicitly by
$$\stackrel{~}{L}\underset{\mu }{}A_\mu (t)_\mu \frac{6(f+g)}{l},$$
(22)
and
$$A_\mu (t)1+\frac{6\tau ^{}}{l\tau }p_\mu (1k(\stackrel{}{e}_\mu ;t)),$$
(23)
where $`_\mu `$ has been defined previously in Eq.(14), $`\mu =\{\pm 1,\pm 2,\pm 3\}`$. It is important to emphasize that all coefficients $`A_\mu (t)=A_\mu (E,V_c;t)`$, i.e. are functions of both the applied field and the carrier velocity.
Now, several comments about equations (20) and (21) are in order. First of all, let us note that Eq.(21) represents, from the mathematical point of view, the boundary conditions for the general evolution equation (20), imposed on the sites in the immediate vicinity of the carrier. Equations (20) and (21) have a different form since in the immediate vicinity of the carrier its asymmetric hopping rules perturb essentially the โenvironmentโ particles dynamics. Equations (20) and (21) possess some intrinsic symmetries and hence the number of independent parameters can be reduced. Namely, reversing the field, i.e. changing $`EE`$, leads to the mere replacement of $`k(\stackrel{}{e}_1;t)`$ by $`k(\stackrel{}{e}_1;t)`$ but does not affect $`k(\stackrel{}{e}_\nu ;t)`$ with $`\nu =\{\pm 2,\pm 3\}`$, which implies that
$$k(\stackrel{}{e}_1;t)(E)=k(\stackrel{}{e}_1;t)(E),\text{and}k(\stackrel{}{e}_\nu ;t)(E)=k(\stackrel{}{e}_\nu ;t)(E)\text{for}\nu =\{\pm 2,\pm 3\},$$
(24)
Besides, since the transition probabilities in Eq.(2) obey
$$p_2=p_2=p_3=p_3$$
(25)
one evidently has that
$$k(\stackrel{}{e}_2;t)=k(\stackrel{}{e}_2;t)=k(\stackrel{}{e}_3;t)=k(\stackrel{}{e}_3;t),$$
(26)
and, by symmetry,
$$A_2(t)=A_2(t)=A_3(t)=A_3(t)$$
(27)
which somewhat simplifies equations (20) and (21). Lastly, we note that despite the fact that using the decoupling scheme in Eq.(19) we effectively close the system of equations on the level of the pair correlations, the solution of Eqs.(20) and (21) still poses serious technical difficulties. Namely, these equations are strongly non-linear with respect to the carrier velocity, which introduces the gradient term on the rhs of the evolution equations for the pair correlation, and depends by itself on the values of the โenvironmentโ particles densities in the immediate vicinity of the carrier. Below we discuss a solution to this non-linear problem, focusing on the limit $`t\mathrm{}`$.
## IV Solution of the decoupled evolution equations in the stationary state.
Consider the limit $`t\mathrm{}`$ and suppose that the density profiles and the stationary velocity of the carrier have non-trivial stationary values
$$k(\stackrel{}{\lambda })\underset{t\mathrm{}}{lim}k(\stackrel{}{\lambda };t),V_c\underset{t\mathrm{}}{lim}V_c(t),\text{and}A_\mu =\underset{t\mathrm{}}{lim}A_\mu (t).$$
(28)
Define next the local deviations of $`k(\stackrel{}{\lambda })`$ from the unperturbed density as
$$h(\stackrel{}{\lambda })k(\stackrel{}{\lambda })\rho _s.$$
(29)
Choosing $`h(\mathrm{๐})=0`$, we obtain the following fundamental system of equations:
$$\stackrel{~}{L}h(\stackrel{}{\lambda })=0,$$
(30)
which holds for $`\stackrel{}{\lambda }\{\mathrm{๐},\stackrel{}{e}_{\pm 1},\stackrel{}{e}_{\pm 2},\stackrel{}{e}_{\pm 3}\}`$, while for the special sites adjacent to the carrier, i.e. for $`\stackrel{}{\lambda }=\{\mathrm{๐},\stackrel{}{e}_{\pm 1},\stackrel{}{e}_{\pm 2},\stackrel{}{e}_{\pm 3}\}`$, one has
$$(\stackrel{~}{L}+A_\nu )h(\stackrel{}{e}_\nu )+\rho _s(A_\nu A_\nu )=0,$$
(31)
Equations (30) and (31) determine the spatial distribution of the deviation from the unperturbed density $`\rho _s`$ in the stationary state. Note also that in virtue of the symmetry relations in Eqs.(26) and (27), $`h(\stackrel{}{e}_{\pm 2})=h(\stackrel{}{e}_{\pm 3})`$ and $`A_2=A_2=A_3=A_3`$.
The method for solving the coupled non-linear Eqs.(8),(30) and (31) is as follows: We first solve these equations supposing that the carrier stationary velocity is a given parameter, or, in other words, assuming that $`A_\nu `$ entering Eqs.(30) and (31) are known. In doing so, we obtain $`h(\lambda )`$ in the parametrized form
$$h(\stackrel{}{\lambda })=h(\stackrel{}{\lambda };A_{\pm 1},A_2).$$
(32)
Then, substituting into Eq.(32) particular values $`\stackrel{}{\lambda }=\{\stackrel{}{e}_{\pm 1},\stackrel{}{e}_{\pm 2},\stackrel{}{e}_{\pm 3}\}`$ and making use of the definition of $`A_\mu `$ in Eq.(23), we find a system of three linear equations with three unknowns of the form
$$A_\nu =1+\frac{6\tau ^{}}{l\tau }p_\nu \left(1\rho _sh(\stackrel{}{e}_\nu ;A_{\pm 1},A_2)\right),$$
(33)
where $`\nu =\{\pm 1,2\}`$, which will allow us to define all $`A_\nu `$ explicitly (and hence, all $`h(\stackrel{}{e}_\nu ))`$. Finally, substituting the results into Eq.(8), which can be written down in terms of $`A_\nu `$ as
$$V_c=\frac{l\sigma }{6\tau ^{}}(A_1A_1),$$
(34)
we arrive at a closed-form equation determining implicitly the stationary velocity.
### A Formal expression for the density profiles in the dynamic percolative environment as seen from the stationary moving carrier.
The general solution of Eqs.(30) and (31) can be most conveniently obtained by introducing the generating function
$$H(w_1,w_2,w_3)\underset{n_1,n_2,n_3}{}h(\stackrel{}{\lambda })w_1^{n_1}w_2^{n_2}w_3^{n_3},$$
(35)
where $`n_1`$,$`n_2`$ and $`n_3`$ are the components of the vector $`\stackrel{}{\lambda }`$, $`\stackrel{}{\lambda }=\stackrel{}{e}_1n_1+\stackrel{}{e}_2n_2+\stackrel{}{e}_3n_3`$. Multiplying both sides of Eqs.(30) and (31) by $`w_1^{n_1}w_2^{n_2}w_3^{n_3}`$ and performing summation, we find then that $`H(w_1,w_2,w_3)`$ is given explicitly by
$$H(w_1,w_2,w_3)=l\frac{_\nu \left(A_\nu (w_{|\nu |}^{\nu /|\nu |}1)h(\stackrel{}{e}_\nu )+\rho _s(A_\nu A_\nu )w_{|\nu |}^\nu \right)}{l_\nu A_\nu (w_{|\nu |}^{\nu /|\nu |}1)6(f+g)},$$
(36)
an expression which allows us to determine the stationary density profiles as seen from the carrier which moves with a constant velocity $`V_c`$.
Inversion of the generating function defined by Eq.(36) yields then, after rather lenghty but straightforward calculations, the following explicit result for the local deviation from the unperturbed density:
$`h(\stackrel{}{\lambda })`$ $`=`$ $`\alpha ^1\{{\displaystyle \underset{\nu }{}}A_\nu h(\stackrel{}{e}_\nu )_\nu `$ (37)
$``$ $`\rho _s(A_1A_1)(_1_1)\}F(\stackrel{}{\lambda }),`$ (38)
where $`F(\stackrel{}{\lambda })`$ is given by
$`F(\stackrel{}{\lambda })=`$ $`\left({\displaystyle \frac{A_1}{A_1}}\right)^{n_1/2}{\displaystyle _0^{\mathrm{}}}e^x\mathrm{I}_{n_1}\left(2{\displaystyle \frac{\sqrt{A_1A_1}}{\alpha }}x\right)\times `$ (39)
$`\times `$ $`\mathrm{I}_{n_2}\left(2{\displaystyle \frac{A_2}{\alpha }}x\right)\mathrm{I}_{n_3}\left(2{\displaystyle \frac{A_2}{\alpha }}x\right)\mathrm{d}x,`$ (40)
and
$`\alpha `$ $`=`$ $`{\displaystyle \underset{\nu }{}}A_\nu +{\displaystyle \frac{6(f+g)}{l}}=`$ (41)
$`=`$ $`A_1+A_1+4A_2+{\displaystyle \frac{6(f+g)}{l}}`$ (42)
Consequently, the particles density distribution as seen from the carrier moving with a constant velocity $`V_c`$ obeys
$`k(\stackrel{}{\lambda })`$ $`=`$ $`\rho _s+\alpha ^1\{{\displaystyle \underset{\nu }{}}A_\nu h(\stackrel{}{e}_\nu _\nu `$ (43)
$``$ $`\rho _s(A_1A_1)(_1_1)\}F(\stackrel{}{\lambda }),`$ (44)
where we have to determine three yet unknown parameters $`A_1`$, $`A_1`$ and $`A_2`$.
To determine these parameters, we set in Eq.(37) $`\stackrel{}{\lambda }=\stackrel{}{e}_1`$, $`\stackrel{}{\lambda }=\stackrel{}{e}_1`$ and $`\stackrel{}{\lambda }=\stackrel{}{e}_2`$, which results in the system of three closed-form equations determining the unknown functions $`A_\nu `$, $`\nu =\{\pm 1,2\}`$,
$$A_\nu =1+\frac{6\tau ^{}}{l\tau }p_\nu \left\{1\rho _s\rho _s(A_1A_1)\frac{det\stackrel{~}{C}_\nu }{det\stackrel{~}{C}}\right\},$$
(45)
where $`\stackrel{~}{C}`$ is a square matrix of the third order defined as
$$\left(\begin{array}{ccc}A_1_1F(\stackrel{}{e}_1)\alpha & A_1_1F(\stackrel{}{e}_1)& A_2_2F(\stackrel{}{e}_1)\\ A_1_1F(\stackrel{}{e}_1)& A_1_1F(\stackrel{}{e}_1)\alpha & A_2_2F(\stackrel{}{e}_1)\\ A_1_1F(\stackrel{}{e}_2)& A_1_1F(\stackrel{}{e}_2)& A_2_2F(\stackrel{}{e}_2)\alpha \end{array}\right)$$
(46)
while $`\stackrel{~}{C}_\nu `$ stands for the matrix obtained from $`\stackrel{~}{C}`$ by replacing the $`\nu `$-th column by a column vector $`\left((_1_1)F(\stackrel{}{e}_\nu )\right)_\nu `$. Equation (43), together with the definition of the coefficients $`A_\nu `$, constitutes the first general result of our analysis defining the density distribution in the percolative environment under study.
### B General force-velocity relation.
Substituting Eqs.(37) and (46) into (34), we find that the stationary velocity of the carrier particle is defined implicitly as the solution of equation:
$$V_c=\frac{\sigma }{\tau }(p_1p_1)(1\rho _s)\left\{1+\rho _s\frac{6\tau ^{}}{l\tau }\frac{p_1det\stackrel{~}{C}_1p_1det\stackrel{~}{C}_1}{det\stackrel{~}{C}}\right\}^1,$$
(47)
where $`\stackrel{~}{C}_1`$ and $`\stackrel{~}{C}_1`$ are the following square matrices of the third order:
$$\stackrel{~}{C}_1=\left(\begin{array}{ccc}(_1_1)F(\stackrel{}{e}_1)& A_1_1F(\stackrel{}{e}_1)& A_2_2F(\stackrel{}{e}_1)\\ (_1_1)F(\stackrel{}{e}_1)& A_1_1F(\stackrel{}{e}_1)\alpha & A_2_2F(\stackrel{}{e}_1)\\ (_1_1)F(\stackrel{}{e}_2)& A_1_1F(\stackrel{}{e}_2)& A_2_2F(\stackrel{}{e}_2)\alpha \end{array}\right)$$
(48)
and
$$\stackrel{~}{C}_1=\left(\begin{array}{ccc}A_1_1F(\stackrel{}{e}_1)\alpha & (_1_1)F(\stackrel{}{e}_1)& A_2_2F(\stackrel{}{e}_1)\\ A_1_1F(\stackrel{}{e}_1)& (_1_1)F(\stackrel{}{e}_1)& A_2_2F(\stackrel{}{e}_1)\\ A_1_1F(\stackrel{}{e}_2)& (_1_1)F(\stackrel{}{e}_2)& A_2_2F(\stackrel{}{e}_2)\alpha \end{array}\right).$$
(49)
Equation (47) represents our second principal result defining the force-velocity relation in the dynamic percolative environment for an arbitrary field and arbitrary rates of the diffusive and renewal processes.
## V Carrier velocity in the limit of small applied field $`E`$, friction coefficient and carrier diffusivity in dynamic percolative environment.
We consider now the case when the applied external field $`E`$ is small. Expanding the transition probabilities $`p_1`$ and $`p_1`$ in the Taylor series up to the first order in powers of the external field, i.e.
$$p_{\pm 1}=\frac{1}{6}\pm \frac{\sigma \beta E}{12}+๐ช\left(E^2\right),$$
(50)
we find that $`V_c`$ defined by Eq.(34) follows
$$V_c\frac{\sigma }{6\tau }\left\{\sigma \beta E(1\rho _s)(h(\stackrel{}{e}_1)h(\stackrel{}{e}_1))\right\}.$$
(51)
On the other hand, Eq.(37) entails that
$$h(\stackrel{}{e}_1)h(\stackrel{}{e}_1)=\frac{2\sigma \rho _s(1\rho _s)\tau ^{}}{l\tau \left(\alpha _0(2A_0/\alpha _0)A_0\right)+2\rho _s\tau }\beta E+๐ช\left(E^2\right),$$
(52)
where
$$A_0=lim_{E0}A_\nu =1+\frac{\tau ^{}}{l\tau }(1\rho _s),$$
(53)
and
$$\alpha _0=lim_{E0}\alpha =6\left(1+\frac{\tau ^{}(1\rho _s)}{l\tau }+\frac{f+g}{l}\right),$$
(54)
while
$`(x)`$ $``$ $`\left\{{\displaystyle _0^{\mathrm{}}}e^t\mathrm{I}_0^2(xt)\left(\mathrm{I}_0(xt)\mathrm{I}_2(xt)\right)dt\right\}^1`$ (55)
$`=`$ $`\left\{P(\mathrm{๐};3x)P(2\stackrel{}{e}_1;3x)\right\}^1,`$ (56)
$`P(\stackrel{}{r};\xi )`$ being the generating function,
$$P(\stackrel{}{r};\xi )\underset{j=0}{\overset{+\mathrm{}}{}}P_j(\stackrel{}{r})\xi ^j,$$
(57)
of the probability $`P_j(\stackrel{}{r})`$ that a walker starting at the origin and performing a Polya random walk on the sites of a three-dimensional cubic lattice will arrive on the $`j`$-th step to the site with the lattice vector $`\stackrel{}{r}`$ .
Consequently, we find that in the limit of a small applied field $`E`$ the force-velocity relation in Eq.(47) attains the physically meaningful form of the Stokes formula $`E=\zeta V_c`$, which signifies that the frictional force exerted on the carrier by the environment particles is $`\mathrm{v}\mathrm{i}\mathrm{s}\mathrm{c}\mathrm{o}\mathrm{u}\mathrm{s}`$. The effective friction coefficient $`\zeta `$ is the sum of two terms,
$$\zeta =\zeta _0+\zeta _{coop}$$
(58)
where the first term represents a mean-field-type result $`\zeta _0=6\tau /\beta \sigma ^2(1\rho _s)`$ (see Eq.(7)), while the second one, $`\zeta _{coop}`$, obeys
$$\zeta _{coop}=\frac{12\rho _s\tau ^{}}{\beta \sigma ^2l(1\rho _s).\left(\alpha _0(2A_0/\alpha _0)A_0\right)}$$
(59)
The second contribution has a more complicated origin and is associated with the cooperative behavior - formation of a inhomogeneous stationary particle distribution around the carrier moving with constant velocity $`V_c`$. Needless to say, such an effect can not be observed within the framework of previous models of dynamic percolation, since there the carrier does not influence the host medium dynamics .
Let us now compare the relative importance of two contributions, i.e. $`\zeta _0`$ and $`\zeta _{coop}`$, to the overall friction. In Fig.2 we plot the ratio $`\zeta /\zeta _0`$ versus the creation rate $`f`$ for three different values of the density $`\rho _s`$, $`\rho _s=0.9,0.7`$ and $`0.5`$, while the annihilation rate is prescribed by the relation $`g=f(1\rho _s)/\rho _s`$. This figure shows that the cooperative behavior clearly dominates at small and moderate $`f`$ (which entails also small values of $`g`$), while for larger $`f`$, when $`\zeta /\zeta _0`$ tends to $`1`$, the mean-field behavior becomes most important. The cooperative behavior also appears to be more pronounced at larger densities $`\rho _s`$.
Consider next some analytical estimates. We start with the situation, in which diffusion of the environment particles is suppressed, i.e. when $`l=0`$. In this case, we get
$$\frac{\zeta _{coop}}{\zeta _0}=\frac{2\rho _s}{(1\rho _s)\left(\frac{2}{y}(y)1\right)},$$
(60)
where
$$y=\frac{1}{3}\left(1+\frac{\tau }{\tau ^{}}\frac{(f+g)}{(1\rho _s)}\right)^1.$$
(61)
Suppose first that $`\rho _s`$ is small, $`\rho _s1`$. Then, $`y1/3(1+\tau /\tau ^{}(f+g))`$ and we can distinguish between two situations: when $`\tau (f+g)/\tau ^{}`$, i.e. when the carrier moves faster than the environment re-organizes itself, and and the opposite limit, $`\tau (f+g)/\tau ^{}`$, when the environment changes very rapidly compared to the motion of the carrier. In the former case we find that $`y1/3`$, which yields $`\zeta _{coop}/\zeta _02\rho _s/(6(1/3)1)`$, $`(1/3)0.7942`$, while in the latter case we have $`y\tau ^{}/3\tau (f+g)`$ and $`\zeta _{coop}/\zeta _0\rho _s\tau ^{}/3\tau (f+g)`$. Note, that in both cases the ratio $`\zeta _{coop}/\zeta _0`$ appears to be small, which signifies that at small densities $`\rho _s`$ the mean-field friction dominates. Such a result is consistent with the behavior depicted in Fig.2 and is not counterintuitive, of course, since in the absence of the particlesโ diffusion, which couples effectively the density evolution at different lattice sites, no significant cooperative behavior can emerge at small densities. On the other hand, at relatively high densities $`\rho _s1`$ and $`\tau /(1\rho _s)\tau ^{}/(f+g)\tau `$, when the carrier moves at much faster rate than the host medium reorganizes itself, we find that $`\zeta _{coop}/\zeta _0\tau ^{}/3\tau (f+g)1`$. This result stems from the circumstance that in sufficiently dense environments modeled by dynamic percolation a highly inhomogeneous density profile emerges even in the absence of particles diffusion; Here, on the one hand, the carrier perturbs significantly the particle density in its immediate vicinity. On the other hand, the density perturbance created by the carrier does not shift the global balance between creation and annihilation events, i.e. the mean particle density still equals $`\rho _s`$, as we set out to show in what follows. The latter constraint induces then appearence of essential correlations in particles distribution and hence, appearence of cooperative behavior.
Let us consider the opposite case when the renewal processes are not allowed, which means that the particles number is conserved and local density in the percolative environment evolves only due to particles diffusion. In this case we find
$$\frac{\zeta _{coop}}{\zeta _0}=\frac{2\tau ^{}\rho _s}{(l\tau +\tau ^{}(1\rho _s))\left(6(1/3)1\right)}$$
(62)
Here, the ratio $`\zeta _{coop}/\zeta _0`$ can be large and the โcooperativeโ friction dominates the mean-field one when $`l\tau \tau ^{}(3\rho _s1)`$, which happens, namely, at sufficiently high densities and in the limit when the carrier moves at a much faster rate than the environment reorganizes itself. Otherwise, the mean-field friction prevails.
To estimate the carrier particle diffusion coefficient $`D_c`$ we assume the validity of the Einstein relation, i.e. $`\beta D_c=\zeta ^1`$. We find that, in the general case, the carrier diffusion coefficient $`D_c`$ reads
$$D_c=\frac{\sigma ^2(1\rho _s)}{6\tau }\left\{1\frac{2\rho _s\tau ^{}}{l\tau }\left(\alpha _0(2A_0/\alpha _0)1+\frac{\tau ^{}(3\rho _s1)}{l\tau }\right)^1\right\}$$
(63)
In the particular case of conserved particles number, when $`f,g0`$ but their ratio $`f/g`$ is kept fixed, $`f/g=\rho _s/(1\rho _s)`$, the latter equation reduces to the classical result
$$D_c^{NK}=\frac{\sigma ^2(1\rho _s)}{6\tau }\left\{1\frac{2\rho _s\tau ^{}}{l\tau }\left(6A_0(1/3)1+\frac{\tau ^{}(3\rho _s1)}{l\tau }\right)^1\right\},$$
(64)
obtained earlier in Refs. and by different analytical techniques. The result in Eq.(64) is known to be exact in the limits $`\rho _s1`$ and $`\rho _s1`$, and serves as a very good approximation for the self-diffusion coefficient in hard-core lattice gases of arbitrary density .
It seems also interesting to analyse how random annihilation and creation of particles can modify the self-diffusion coefficient compared to the situation when the particles number is conserved. In Figure 3 we plot the ratio $`D_c^{NK}/D_c`$ ($`=\zeta /\zeta _{NK}`$) versus the creation rate $`f`$ for three different values of the density $`\rho _s`$, $`\rho _s=0.9,0.7`$ and $`0.5`$. Again, the value of the annihilation rate $`g`$ is prescribed by the relation $`g=f(1\rho _s)/\rho _s`$. Figure 3 shows that the renewal processes affect considerably the friction coefficient and the ratio $`\zeta /\zeta _{NK}`$ deviates strongly from the unity with the growth of the creation rate. The overall friction also falls off when the density increases.
Finally, in the absence of particle diffusion (fluctuating-site percolation), our result for the carrier particle diffusion coefficient reduces to
$$D_c^{per}=\frac{\sigma ^2(1\rho _s)}{6\tau }\left\{12\rho _s\left(4[(1\rho _s)+(f+g)\tau /\tau ^{}](y)+3\rho _s1\right)^1\right\}$$
(65)
Note, however, that this result only applies when both $`f`$ and $`g`$ are larger than zero, such that the renewal processes take place. In fact, the underlying decoupling scheme is only plausible in this case. Similarly to the approximate theories in Refs. and , our approach predicts that in the absence of the renewal processes $`D_c^{per}`$ vanishes only when $`\rho _s1`$, which is an incorrect behavior.
## VI Asymptotic behavior of the density profiles at large distances in front of and past the carrier.
The density profiles at large separations in front of and past the carrier can be readily deduced from the asymptotical behavior of the following generating function
$$N(w_1)\underset{n_1=\mathrm{}}{\overset{+\mathrm{}}{}}h(n_1,n_2=0,n_3=0)w_1^{n_1}.$$
(66)
Inversion of Eq.(36) with respect to the symmetric coordinates $`n_2`$ and $`n_3`$ yields then
$`N(w_1)={\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)+\rho _s(A_1A_1)\right)\left(w_11\right)+\left(A_1h(\stackrel{}{e}_1)\rho _s(A_1A_1)\right)\left(w_1^11\right)}{\alpha A_1w_1^1A_1w_1}}\times `$ (67)
$`\times `$ $`{\displaystyle _0^{\mathrm{}}}\mathrm{exp}[x]\mathrm{I}_0^2({\displaystyle \frac{2A_2}{\alpha A_1w_1^1A_1w_1}}x)\mathrm{d}x+{\displaystyle \frac{4A_2h(\stackrel{}{e}_2)}{\alpha A_1w_1^1A_1w_1}}\times `$ (68)
$`\times `$ $`{\displaystyle _0^{\mathrm{}}}\mathrm{exp}[x]\mathrm{I}_0({\displaystyle \frac{2A_2}{\alpha A_1w_1^1A_1w_1}}x)(\mathrm{I}_1({\displaystyle \frac{2A_2}{\alpha A_1w_1^1A_1w_1}}x)`$ (69)
$``$ $`\mathrm{I}_0({\displaystyle \frac{2A_2}{\alpha A_1w_1^1A_1w_1}}x))\mathrm{d}x`$ (70)
We notice now that $`N(w_1)`$ is a holomorphic function in the region $`๐ฒ_1<w_1<๐ฒ_2`$, where
$`๐ฒ_1={\displaystyle \frac{\alpha 4A_2}{2A_1}}\sqrt{\left({\displaystyle \frac{\alpha 4A_2}{2A_1}}\right)^2{\displaystyle \frac{A_1}{A_1}}}`$ (71)
and
$`๐ฒ_2={\displaystyle \frac{\alpha 4A_2}{2A_1}}+\sqrt{\left({\displaystyle \frac{\alpha 4A_2}{2A_1}}\right)^2{\displaystyle \frac{A_1}{A_1}}}`$ (72)
As a consequence, the asymptotic behavior of $`h(n_1,n_2=0,n_3=0)`$ in the limit $`n_1\mathrm{}`$ (resp. $`n_1\mathrm{}`$) is controlled by the behavior of $`N(w_1)`$ in the vicinity of $`w_1=๐ฒ_2`$ (resp. $`w_1=๐ฒ_1`$) (see, for example, the analysis of the generating function singularities developed in Ref.).
#### 1 Asymptotics of the density profiles at large separations in front of the carrier.
Consider first the asymptotic behavior of the density distribution of the โenvironmentโ particles at large separations in front of the carrier. Using the fact that
$$_0^{\mathrm{}}exp\left[x\right]\mathrm{I}_0(yx)\left(\mathrm{I}_1(x)\mathrm{I}_0(x)\right)dx$$
(73)
is a regular function when $`y1/2`$, while
$$_0^{\mathrm{}}exp\left[x\right]\mathrm{I}_0^2(yx)dx\frac{1}{\pi }\mathrm{ln}\left(\frac{1}{12y}\right),$$
(74)
we find that
$`N(w_1)`$ $``$ $`{}_{w_1๐ฒ_2}{}^{}[{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)+\rho _s(A_1A_1)\right)\left(๐ฒ_21\right)}{4\pi A_2}}+`$ (75)
$`+`$ $`{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)\rho _s(A_1A_1)\right)\left(๐ฒ_2^11\right)}{4\pi A_2}}\left]\mathrm{ln}\right(๐ฒ_2w_1)`$ (76)
Then, (cf, Flajolet et al., Ref.), we obtain the following asymptotical result
$$h(n_1,0,0)_{n_1\mathrm{}}\frac{K^+}{n_1}e^{n_1/\lambda _+},$$
(77)
where
$$\lambda _+\mathrm{ln}^1\left(\frac{\alpha /22A_2}{A_1}+\sqrt{\left(\frac{\alpha /22A_2}{A_1}\right)^2\frac{A_1}{A_1}}\right),$$
(78)
and
$`K^+`$ $`=`$ $`[{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)+\rho _s(A_1A_1)\right)\left(๐ฒ_21\right)}{4\pi A_2}}+`$ (79)
$`+`$ $`{\displaystyle \frac{(A_1h(\stackrel{}{e}_1)\rho _s(A_1A_1)(๐ฒ_2^11)}{4\pi A_2}}]>0,`$ (80)
which signifies that the density of the โenvironmentโ particles in front of the carrier is higher than the average value $`\rho _s`$ and approaches $`\rho _s`$ at large separations from the carrier as an exponential function of the distance.
#### 2 Asymptotics of the density profiles at large separations past the carrier.
We consider next the asymptotic behavior of the โenvironmentโ particles density profiles past the carrier particle, which turns out to be very different depending on whether the dynamics of the percolative environment obeys the strict conservation of the โenvironmentโ particles number or not (the renewal processes are suppressed or allowed). The sketch of this behavior is presented in Fig.4.
##### a Non-conserved particles number.
In the case when partciles may disappear and re-appear on the lattice, one has that the root $`๐ฒ_1<1`$. We find then, following essentially the same lines as in the previous subsection, that
$`N(w_1)`$ $``$ $`{}_{w_1๐ฒ_1}{}^{}[{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)+\rho _s(A_1A_1)\right)\left(๐ฒ_11\right)}{4\pi A_2}}+`$ (81)
$`+`$ $`{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)\rho _s(A_1A_1)\right)\left(๐ฒ_1^11\right)}{4\pi A_2}}\left]\mathrm{ln}\right({\displaystyle \frac{1}{w_1๐ฒ_1}}).`$ (82)
Hence, in the non-conserved case the approach to the unperturbed value $`\rho _s`$ is also exponential when $`n_1\mathrm{}`$, and follows
$$h_{n_1,0,0}_{n_1\mathrm{}}\frac{K^{}}{|n_1|}e^{|n_1|/\lambda _{}},$$
(83)
where
$$\lambda _{}\mathrm{ln}^1\left(\frac{\alpha /22A_2}{A_1}\sqrt{\left(\frac{\alpha /22A_2}{A_1}\right)^2\frac{A_1}{A_1}}\right)$$
(84)
and
$`K^{}`$ $`=`$ $`[{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)+\rho _s(A_1A_1)\right)\left(๐ฒ_11\right)}{4\pi A_2}}+`$ (85)
$`+`$ $`{\displaystyle \frac{\left(A_1h(\stackrel{}{e}_1)\rho _s(A_1A_1)\right)\left(๐ฒ_1^11\right)}{4\pi A_2}}]<0`$ (86)
which implies that the particles density past the carrier is lower than the average. Note that, in the general case, $`\lambda _+<\lambda _{}`$, which means that the depleted region past the carrier is more extended in space than the traffic-jam-like region in front of the carrier. The density profiles are therefore asymmetric with respect to the origin, $`n_1=0`$. Since creation of particles is favored (suppressed) in depleted (jammed) regions, while annihilation is suppressed (favored), one might expect that this will shift the overall density in the system, i.e. the average density of the โenvironmentโ particles will differ from $`\rho _s`$. Interestingly, the overall deviation, i.e. the sum of local deviations over the volume of the system, of the density of the โenvironmentโ particles from the average value $`\rho _s`$, appears to be equal exactly to zero,
$$H(w_1=1,w_2=1,w_3=1)0,$$
(87)
and hence, the driven carrier does not perturb the global balance between creation and annihilation of the โenvironmentโ particles. This is not, however, an $`\mathrm{a}\mathrm{p}\mathrm{r}\mathrm{i}\mathrm{o}\mathrm{r}\mathrm{i}`$ evident result in view of the asymmetry of the density profiles.
##### b Conserved particles number.
Finally, we turn to the analysis of the shape of the density profiles of the percolative environment past the carrier in the particular limit when the host medium evolves only due to diffusion, while creation and annihilation of particles are completely suppressed. In this case, in which the particles number is explicitly conserved, one has that for arbitrary value of the field and particlesโ average density, the root $`๐ฒ_11`$ and, consequently, the form of the generating function is qualitatively different from that in Eqs.(75) and (81),
$`N(w_1)`$ $``$ $`{}_{w_11^+}{}^{}[{\displaystyle \frac{(A_1h(\stackrel{}{e}_1)A_1h(\stackrel{}{e}_1)}{4\pi A_2}}+`$ (88)
$`+`$ $`{\displaystyle \frac{2\rho _s(A_1A_1))}{4\pi A_2}}](w_11)\mathrm{ln}({\displaystyle \frac{1}{w_11}}).`$ (89)
Equation (88) implies that in the limit when the particle number is conserved the large-$`n_1`$ asymptotic behavior of $`h_{n_1,0,0}`$ is described by an algebraic function of $`n_1`$ with a logarithmic correction; that is,
$$h_{n_1,0,0}\frac{K_{}\mathrm{ln}(|n_1|)}{n_1^2},$$
(90)
where $`K_{}`$ is an $`n_1`$-independent constant. Remarkably, the power-law decay of correlations implies existence of a quasi-long-range order in the percolative environment past the carrier. In the conserved case the mixing of the three-dimensional percolative environment is not very efficient and there are considerable memory effects - the host medium remembers the passage of the carrier on large space and time scales.
## VII Conclusions
To conclude, we have studied analytically the dynamics of a carrier driven by an external field $`\stackrel{}{E}`$ in a three-dimensional environment modeled by dynamic percolation on cubic lattice partially filled with mobile, hard-core โenvironmentโ particles which can spontaneously disappear and reappear (renewal processes) in the system with some prescribed rates. Our analytical approach has been based on the master equation, describing the time evolution of the system, which has allowed us to evaluate a system of coupled dynamical equations for the carrier velocity and a hierarchy of correlation functions. To solve these coupled equations, we have invoked an approximate closure scheme based on the decomposition of the third-order correlation functions into a product of pairwise correlations, which has been first introduced in Ref. for a related model of a driven carrier dynamics in a one-dimensional lattice gas with conserved particles number. Within the framework of this approximation, we have derived a system of coupled, discrete-space equations describing evolution of the density profiles of the environment, as seen from the moving carrier, and its velocity $`V_c`$. We have shown that $`V_c`$ depends on the density of the โenvironmentโ particles in front of and past the carrier. Both densities depend on the magnitude of the velocity, as well as on the rate of the renewal and diffusive processes. As a consequence of such a non-linear coupling, in the general case, (i.e. for an arbitrary driving field and arbitrary rates of renewal and diffusive processes), $`V_c`$ has been found only implicitly, as the solution of a non-linear equation relating its value to the system parameters. This equation, which defines the force-velocity relation for the dynamic percolation under study, simplifies considerably in the limit of small applied field $`\stackrel{}{E}`$. We find that in this limit it attains the physically meaningful form of the Stokes formula, which implies, in particular, that the frictional force exerted on the carrier by the environment modeled by dynamic percolation is $`viscous`$. In this limit, the carrier velocity and the friction coefficient are calculated explicitly. In addition, we determine the self-diffusion coefficient of the carrier in the absence of the field and show that it reduces to the well-know result of Refs. and in the limit when the particles number is conserved. Further more, we have found that the density profile around the carrier becomes strongly inhomogeneous: the local density of the โenvironmentโ particles in front of the carrier is higher than the average and approaches the average value as an exponential function of the distance from the carrier. On the other hand, past the carrier the local density is lower than the average, and depending on whether the number of particles is explicitly conserved or not, the local density past the carrier may tend to the average value either as an exponential or even as an $`\mathrm{a}\mathrm{l}\mathrm{g}\mathrm{e}\mathrm{b}\mathrm{r}\mathrm{a}\mathrm{i}\mathrm{c}`$ function of the distance. The latter reveals especially strong memory effects and strong correlations between the particle distribution in the environment and the carrier position.
Figure Captions.
Fig.1. A generalized model of dynamic percolation. Grey spheres denote the hard-core โenvironmentโ particles, which perform symmetric random hopping among the sites of a simple cubic lattice, and can be spontaneously annihilated and created. The lighter sphere is the carrier, which performs a biased random walk due to an external field $`\stackrel{}{E}`$, consrained by hard-core exclusion with the โenvironmentโ particles.
Fig.2. The ratio of the overall friction coefficient and the mean-field friction versus the creation rate for three different values of the mean density $`\rho _s`$. The upper curve corresponds to $`\rho _s=0.9`$, the intermediate - to $`\rho _s=0.7`$, and the lower - to $`\rho _s=0.5`$.
Fig.3. The ratio of the overall friction coefficient and the friction coefficient in the conserved particles number case versus the creation rate for three different values of the mean density $`\rho _s`$. The upper curve corresponds to $`\rho _s=0.5`$, the intermediate - to $`\rho _s=0.7`$, and the lower - to $`\rho _s=0.9`$.
Fig.4. A sketch of the asymptotic density profiles in front of and past the stationary moving carrier. The abbreviation CPN stands for the โconserved particles numberโ. The two solid lines in the $`\lambda >0`$ and $`\lambda <0`$ domains denote exponential profiles, Eqs.(63) and (67). The dashed line in the domain $`\lambda <0`$ stands for the algebraic law, Eq.(72). |
warning/0003/hep-ph0003288.html | ar5iv | text | # Flow equations in the light-front QCD: mass gap and confinement
## 1 Introduction
The perturbative aspects of non-abelian gauge theories were underestood many years ago, and the perturbative calculations provided convincing proof that QCD is the theory of strong interactions. However nonperturbative QCD phenomena have been difficult to analyze mainly because calculational techniques are still lacking, even though the qualitative features have been more or less understood.
In particular, it is widely believed that pure Yang-Mills theory, with no dynamical quarks, posseses the features like asymptotic freedom, mass generation through the transmutation of dimensions and confinement: linear potential between static (probe) quarks. In the past few years there were several studies addressing the issue of confinement and generation of mass gap in the framework of the Schrรถdinger picture $`^\mathrm{?}`$, $`^\mathrm{?}`$. To mention a few early works, see refs. $`^\mathrm{?}`$. We have tried to understand these nonperturbative mechanisms also in a Hamiltonian framework, solving flow equations for canonical QCD Hamiltonian in the light-front (LF) gauge selfconsistently for the few lowest sectors. Dynamics of quarks has been excluded to disentangle the complexity of chiral symmetry breaking. Early studies of confinement in the LF frame were conducted using similarity renormalization $`^\mathrm{?}`$ and transverse lattice $`^\mathrm{?}`$, based on the fact that QCD in $`3+1`$ already has a confining interaction in the form of the instantaneous four-fermion interaction, which is the confining interaction in $`1+1`$. However, in our study the instantaneous interaction is canceled by the corresponding term in the dynamical interaction, generated by flow equations. The result is a three dimensional linear rising potential in the infrared region.
Our basic strategy has been to generate dynamically an effective gluon mass (energy) through interaction with the LF โvacuumโ (exact zero modes on the light-front, $`x=0`$), and construct an effective interaction between static quarks as an exchange with this dynamical gluon, which forms a flux between quarks. One could construct a LF field theory with massive photons and gluons, as was done in $`^\mathrm{?}`$, but this theory is in a Coulomb phase rather than confining because of the lack of a dynamical mass generation mechanism. Dynamical mass generation in the light-front frame dates back to the works of Cornwall $`^\mathrm{?}`$, where a kind of pinch technique was used to define a gluon propagator and hence a gluon effective energy, extracted from a Greenโs function for physical observables (for example two-body scattering amplitude). It was suggested in $`^\mathrm{?}`$, that in physical terms, a gluon โmassโ may lead to a vortex condensation <sup>a</sup><sup>a</sup>aIn the massive gauge theory one needs to introduce a scalar field, which condenses, in order to preserve gauge invariance. Topologically this field represents vortices $`^\mathrm{?}`$. with $`<TrG_{\mu \nu }^2>0`$, and conversely $`<TrG_{\mu \nu }^2>0`$ generates a gluon mass. This provides a link with the existing vortex picture of confinement.
A subtle point in the LF field theory is that the LF vacuum is just empty space. Therefore it seems a problem how confinement can occur in the LF frame, and what quantity sets up a scale for a dynamical mass gap and the string tension. We adopt a model, suggested by Susskind and Burkardt in the context of chiral symmetry breaking in the LF frame $`^\mathrm{?}`$. In the parton model one pictures a fast moving hadron as being some collection of constituents with relatively large momentum, such that when one boosts the system, doubles its momentum, all these partons double their momenta and so forth. Therefore one can formulate a field theory on the axis of the light-front momentum $`x`$. Partons that form a hadron are at positive, finite $`x`$ and according to Feynman and Bjorken fill the $`x`$-axis in a way which gets denser and denser as one goes to smaller $`x`$; and the vacuum is at $`x=0`$. The fundamental property of LF Hamiltonians, that under a rescaling of the LF momentum, $`x\lambda x`$, the LF Hamiltonian scales like $`HH/\lambda `$, can be interpreted as a dilatation symmetry along the $`x`$-axis, if one thinks of the $`x`$-axis as a spacial axis. This symmetry holds on a classical level and it is brocken on a quantum level by anomali. As one approaches small $`x`$, interaction between partons gets stronger, contributing divergent matrix elements. A natural cutoff is provided by $`\delta x=\epsilon x`$, a minimal spacing between constituents, which plays the role of UV-regulator <sup>b</sup><sup>b</sup>bSmall $`x`$ correspond to the large light-front energies.. As long as $`\delta x`$ is finite, i.e. as long as the density of partons on the $`x`$-axis is not infinite, one obtains finite matrix elements. Cutoff $`\delta x`$ breaks the dilatation symmetry along the $`x`$-axis and gauge symmetry, and sets up a scale for quantities of dimension of energy (mass) and higher powers in energy. In particular, one generates dynamically an effective gluon mass, which in turn will define string tension between quark and antiquark. Formation of the $`q\overline{q}`$ bound state through breaking an internal symmetry can be viewed analogously to the creation of Cooper pairs in superconducter.
In terms of effective theory, the Hamiltonian below the LF cutoff, $`H_\epsilon `$, which desribes high energy (UV) strongly correlated states, can be substituted by the v.e.v., since strongly coupled configurations are frozen. Hamiltonian above the cutoff $`H_{>\epsilon }`$ is treated by flow equations (renormalization group transformation). One way of looking at the physics behind this v.e.v. and mass generation in gluodynamics is that the composite field $`\varphi `$ which creates $`0^+`$ glueballs has a finite v.e.v. $`^\mathrm{?}`$, $`^\mathrm{?}`$.
The dilatation symmetry, discussed above, reflects some underlying scale invariance of the LF field theory formulated on $`x`$-axis. Introducing the cutoff, breaks this symmetry. Because physics should remain independent on the cutoff, one must be looking for a fixed point of the renormalization group. Therefore the right tool for studing such a system is the renormalization group, which is provided by the method of flow equations $`^\mathrm{?}`$.
In the method of flow equations an effective interaction between quarks arises through elimination of quark-gluon coupling. Procedure converges in the UV for large gluon momenta transfer $`\stackrel{}{q}^2`$, but in the IR one is not able to eliminate vanishing gluon momenta, because the similarity factor which governes the elimination does not decay for vanishing arguments. In physical terms, one can not integrate soft gluons in the same fashion. As soon as gluon aquires a dynamical mass, which vanishes at large gluon momenta and not equal zero only at small momenta, the similarity factor decays even for small momenta transfer, because the argument contains an effective gluon energy instead of only gluon mometum. Therefore we can eliminate by flow equations even soft gluon modes, that are responsible for the long-range part of an effective quark potential. In the IR this gives $`1/q^4`$ behavior for an effective $`q\overline{q}`$ interaction.
In the next section we solve flow equations for an effective gluon mass and quark-antiquark effective interaction, based on the QCD Hamiltonian in the light-front gauge.
## 2 Gluon mass gap and an effective quark interaction
Let $`Q`$ being a projection operator on a one-gluon state, and $`P`$ on a $`q\overline{q}`$ state, $`Q|\psi =|g`$ and $`P|\psi =|q\overline{q}`$. Flow equations for matrix elements of the QCD Hamiltonian between these states read
$`{\displaystyle \frac{dE_q(l)}{dl}}`$ $`=`$ $`{\displaystyle \underset{p}{}}{\displaystyle \frac{1}{E_q(l)E_p(l)}}{\displaystyle \frac{d}{dl}}\left(h_{qp}(l)h_{pq}(l)\right)`$ (1)
$`{\displaystyle \frac{dh_{pp^{}}(l)}{dl}}`$ $`=`$ $`{\displaystyle \underset{q}{}}\left({\displaystyle \frac{dh_{pq}(l)}{dl}}{\displaystyle \frac{1}{E_p(l)E_q(l)}}h_{qp^{}}(l)+h_{pq}(l){\displaystyle \frac{1}{E_p^{}(l)E_q(l)}}{\displaystyle \frac{dh_{qp^{}}(l)}{dl}}\right),`$
where $`p`$ ($`p^{}`$) runs through all states in the $`P`$-subspace, and $`q`$ in the $`Q`$-subspace. Flow equations for the quark-gluon coupling $`h_{pq}`$ (coupling between $`P`$ and $`Q`$ sectors) are
$`{\displaystyle \frac{dh_{pq}(l)}{dl}}`$ $`=`$ $`\left(E_p(l)E_q(l)\right)\eta _{pq}(l)`$
$`\eta _{pq}(l)`$ $`=`$ $`{\displaystyle \frac{h_{pq}(l)}{E_p(l)E_q(l)}}{\displaystyle \frac{d}{dl}}\left(\mathrm{ln}f(z_{pq}(l))\right)`$
$`z_{pq}(l)`$ $`=`$ $`l\left(E_p(l)E_q(l)\right)^2,`$ (2)
where $`f`$ is the similarity function with properties $`f(0)=1`$ and $`f(z\mathrm{})=0`$. From Eq. (2), elimination of quark-gluon coupling
$`h_{pq}(l)=h_{pq}(0)f(z_{pq}(l)),`$ (3)
is governed by the similarity function $`f(l(E_p(l)E_q(l))^2)`$, which vanishes for the matrix elements with energy differences exceeding the cutoff $`\lambda `$, $`|E_p(l)E_q(l)|1/\sqrt{l}=\lambda `$, $`h_{pq}(0)`$ is the initial value, and $`E_q(l)`$ ($`E_p(l)`$) is a solution of the flow (renormalization group) equation, Eq. (1). As long as the argument of the similarity function is not equal zero, one can eliminate quark-gluon coupling and solve flow equations for an effective gluon energy and effective $`q\overline{q}`$-interaction, Eq. (1). However, if one neglects the cutoff dependence of energies, $`E_p(l)E_p(0)`$, the argument of the similarity function may become zero in the degenerate case, $`E_p(0)=E_q(0)`$, and then the energy denominator blows up in effective interactions in Eq. (1).
In physical terms, the leading behavior of the argument in the quark-gluon similarity function is given by an effective gluon energy transfer between quark and antiquark, $`\left(E_p(l)E_q(l)\right)E_{gluon}^{eff}(l)`$, which reduces at the initial value of flow parameter, $`l=0`$, to a gluon momentum transfer, $`E_{gluon}^{eff}(0)\stackrel{}{q}^2`$. Neglecting the cutoff dependence of the effective gluon energy, one eliminates quark-gluon coupling only at large gluon momenta, and fails at small gluon momenta, because the argument of the similarity function tends to zero. For small gluon momenta the dependence of gluon effective energy on the cutoff becomes important, given by a dynamically generated gluon mass, that makes elimination of quark-gluon coupling possible even for small gluon momenta. We therefore solve flow equations, Eq. (1), consistently for an effective gluon mass and an effective $`q\overline{q}`$-interaction below. We show, that elimination of quark-gluon coupling for high and low gluon momenta contribute correspondingly to the UV and IR parts of quark-antiquark potential, that behave in momentum space like $`1/\stackrel{}{q}^2`$ and $`1/\stackrel{}{q}^4`$, respectively, where $`\stackrel{}{q}`$ is a gluon momentum transfer.
### 2.1 Gluon gap equation
In the light-front frame flow equation for an effective gluon mass (the first equation in Eq. (1)), with connection to the light-front energy $`q^{}=(q_{}^2+\mu ^2(\lambda ))/q^+`$, reads
$`{\displaystyle \frac{d\mu ^2(\lambda )}{d\lambda }}`$ $`=`$ $`2T_fN_f{\displaystyle _0^1}{\displaystyle \frac{dx}{x(1x)}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d^2k_{}}{16\pi ^3}}g_q^2(\lambda ){\displaystyle \frac{1}{Q_2^2(\lambda )}}{\displaystyle \frac{df^2(Q_2^2(\lambda );\lambda )}{d\lambda }}`$ (4)
$`\times `$ $`\left({\displaystyle \frac{k_{}^2+m^2}{x(1x)}}2k_{}^2\right)`$
$`+`$ $`2C_a{\displaystyle _0^1}{\displaystyle \frac{dx}{x(1x)}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d^2k_{}}{16\pi ^3}}g_g^2(\lambda ){\displaystyle \frac{1}{Q_1^2(\lambda )}}{\displaystyle \frac{df^2(Q_1^2(\lambda );\lambda )}{d\lambda }}`$
$`\times `$ $`\left(k_{}^2(1+{\displaystyle \frac{1}{x^2}}+{\displaystyle \frac{1}{(1x)^2}})\right),`$
with
$`Q_1^2(\lambda )`$ $`=`$ $`{\displaystyle \frac{k_{}^2+\mu ^2(\lambda )}{x(1x)}}\mu ^2(\lambda ),Q_2^2(\lambda )={\displaystyle \frac{k_{}^2+m^2}{x(1x)}}\mu ^2(\lambda ),`$ (5)
where gluon couples to the quark-anti-quark pairs, with the quark-gluon coupling $`g_q(\lambda )`$, and pairs of gluons, with the three-gluon coupling $`g_g(\lambda )`$; for finite $`\lambda `$ these couplings are different from each other functions of momenta; connection between flow parameter and the cutoff, $`l=1/\lambda ^2`$, is used; similarity function $`f`$ plays the role of UV regulator in loop integrals; current quark mass is $`m`$; for $`SU(N_c)`$ $`T_f\delta _{ab}=\mathrm{Tr}(T^aT^b)=\frac{1}{2}\delta _{ab}`$ and the adjoint Casimir is $`C_a\delta _{ab}=f^{acd}f^{bcd}=N_c\delta _{ab}`$, $`N_c`$ is the number of colors (i.e., $`N_c=3`$).
Coupled system of equations for the cutoff dependent, effective gluon and quark masses was first derived in $`^\mathrm{?}`$. Here, neglecting the cutoff dependence of quark mass this system is reduced to Eq. (4).
Generally, it is difficult to solve Eq. (4). One of the reasons is that this equation contains (unknown) running coulings; therefore the cutoff dependence of couplings is neglected below. Second, initial condition for Eq. (4) is not known. The following renormalization condition to define the running gluon mass $`\mu (\lambda )`$ through the โphysicalโ mass is assumed $`^\mathrm{?}`$: the effective Hamiltonian at the scale $`\lambda `$ has bosonic eigenstates with eigenvalues of the form $`q^{}=(q_{}^2+\stackrel{~}{\mu }^2)/q^+`$ $`^\mathrm{?}`$, i.e.
$`{\displaystyle \frac{q_{}^2+\stackrel{~}{\mu }^2}{q^+}}q^{},q`$ $`=`$ $`{\displaystyle \frac{q_{}^2+\mu ^2(\lambda )}{q^+}}q^{},q`$ (6)
$``$ $`{\displaystyle ^\lambda }๐\lambda ^{}{\displaystyle \underset{p}{}}\left(\eta _{q^{}p}(\lambda ^{})h_{pq}(\lambda ^{})h_{q^{}p}(\lambda ^{})\eta _{pq}(\lambda ^{})\right),`$
where $`\stackrel{~}{\mu }`$ denotes the โphysicalโ gluon mass; the generator $`\eta _{pq}`$, given by Eq. (2), eliminates the quark-gluon (three-gluon) coupling $`h_{pq}`$; $`|q`$ denotes a single effective gluon state with momentum $`q^+`$ and $`q_{}`$, $`q^{}|q=16\pi ^3q^+\delta ^{(3)}(q^{}q)`$. Neglecting the cutoff dependence of the gluon mass on the r.h.s. of Eq. (4), the renormalization point is choisen at $`q^2=\stackrel{~}{\mu }^2`$ (for in- and out-going and intermediate state gluons). Third, severe divergences arise at small light-front momenta $`x`$ and must be regularized. For this purpose, as suggested by Zhang and Harindranath in $`^\mathrm{?}`$, the minimum cutoff $`u`$ for transverse momentum $`k_{}`$ is introduced. Therefore, flow equation, Eq. (4), is integrated in the finite limits, $`[u;\lambda ]`$. Explicitly, from the similarity function in gluon loop (the second term in Eq. (4)) one has $`u\stackrel{~}{Q}_1^2=(k_{}^2+\stackrel{~}{\mu }^2)/x(1x)\stackrel{~}{\mu }^2\lambda `$, that restricts the transverse momentum to $`k_{min}=(u^2+\stackrel{~}{\mu }^2)x(1x)\stackrel{~}{\mu }^2k_{}k_{max}=(\lambda ^2+\stackrel{~}{\mu }^2)x(1x)\stackrel{~}{\mu }^2`$, and the light-front momentum to $`\stackrel{~}{\mu }^2/(u^2+\stackrel{~}{\mu }^2)x1\stackrel{~}{\mu }^2/(u^2+\stackrel{~}{\mu }^2)`$. In the same way one finds restrictions on mometa in the quark-gluon loop (the first term in Eq. (4)). Integrating Eq. (4) with all constraints discussed above and assuming the condition of Eq. (6), gives $`^\mathrm{?}`$
$`\mu ^2(\lambda )`$ $`=`$ $`\stackrel{~}{\mu }^2+\delta \mu _{PT}^2(\lambda )+\delta \mu _{dyn}^2(\lambda ,\stackrel{~}{\mu },u).`$ (7)
where the perturbative term
$`\delta \mu _{PT}^2(\lambda )={\displaystyle \frac{g^2}{4\pi ^2}}\lambda ^2\left(C_a(\mathrm{ln}{\displaystyle \frac{u^2}{\stackrel{~}{\mu }^2}}{\displaystyle \frac{11}{12}})+T_fN_f{\displaystyle \frac{1}{3}}\right).`$ (8)
reproduces the result of the LF perturbation theory, with renormalization point $`q^2=0`$, $`^\mathrm{?}`$. When renormalization is performed in the second oder, the perturbative mass correction is absorbed by the mass counterterm, $`m_{CT}^2=\delta \mu _{PT}^2(\lambda =\mathrm{\Lambda }\mathrm{})`$. After the perturbative renormalization is completed, the dynamical mass
$`\mu _{dyn}^2(\lambda )`$ $`=`$ $`\stackrel{~}{\mu }^2+\delta \mu _{dyn}^2(\lambda ,\stackrel{~}{\mu },u)=\stackrel{~}{\mu }^2+\sigma (\stackrel{~}{\mu },u)\mathrm{ln}{\displaystyle \frac{\lambda ^2}{\stackrel{~}{\mu }^2}}`$
$`\sigma (\stackrel{~}{\mu },u)`$ $`=`$ $`{\displaystyle \frac{g^2}{4\pi ^2}}\stackrel{~}{\mu }^2\left(C_a({\displaystyle \frac{u^2}{\stackrel{~}{\mu }^2}}+\mathrm{ln}{\displaystyle \frac{u^2}{\stackrel{~}{\mu }^2}}{\displaystyle \frac{5}{12}})+T_fN_f({\displaystyle \frac{1}{3}}+{\displaystyle \frac{m^2}{\stackrel{~}{\mu }^2}})\right),`$ (9)
is left. Note, that the gluon mass $`\mu _{dyn}`$ is generated dynamically by flow equations. In the limit $`\stackrel{~}{\mu }0`$, one has $`\sigma =lim_{\stackrel{~}{\mu }0}\sigma (\stackrel{~}{\mu },u)=u^2g^2C_a/2\pi ^2`$, and, as shown below, $`\sigma `$ plays the role of the string tension between quark and antiquark. Scale $`u`$ is introduced by regulating the divergences at small LF momenta, $`x0`$, (small LF $`x`$ regularization). The value $`u`$ sets up a scale for the dynamical gluon mass and the string tension.
### 2.2 Effective $`q\overline{q}`$-interaction
An effective interaction, generated between quark and antiquark by flow equations in the LF frame (the second equation in Eq. (1)), reads
$`V_{q\overline{q}}=Const\gamma ^\mu \gamma ^\nu \underset{\stackrel{~}{\mu }0}{lim}B_{\mu \nu },`$ (10)
where instead of the strong coupling $`\alpha _s`$ some constant, $`Const`$, is introduced; $`\gamma ^\mu \gamma ^\nu `$ is a current-current term in the exchange channel. At the end we set the gluon mass renormalization point (โphysicalโ gluon mass) to zero, $`\stackrel{~}{\mu }0`$; thus eliminating the sensitivity to the renormalization point. The kernel includes dynamical interaction and instantaneous exchange with nonzero dynamical gluon mass, given by
$`B_{\mu \nu }=g_{\mu \nu }\left(I_1+I_2\right)+\eta _\mu \eta _\nu {\displaystyle \frac{\delta Q^2}{q^{+2}}}\left(I_1I_2\right),`$ (11)
where $`g_{\mu \nu }`$ is the LF metric tensor, and the LF vector $`\eta _\mu `$ is defined as $`\eta k=k^+`$. The cutoff dependence of four-momentum transfers along quark and antiquark lines is accumulated in the factor
$`I_1={\displaystyle _0^{\mathrm{}}}๐\lambda {\displaystyle \frac{1}{Q_1^2(\lambda )}}{\displaystyle \frac{df(Q_1^2(\lambda );\lambda )}{d\lambda }}f(Q_2^2(\lambda );\lambda ),`$ (12)
with mometum transfers defined in the light-front frame as
$`Q_1^2(\lambda )`$ $`=`$ $`Q_1^2+\mu _{dyn}^2(\lambda ),Q_1^2={\displaystyle \frac{(x^{}k_{}xk_{}^{})^2+m^2(xx^{})^2}{xx^{}}}`$
$`Q_2^2(\lambda )`$ $`=`$ $`Q_2^2+\mu _{dyn}^2(\lambda ),Q_2^2=Q_1^2|_{x(1x);x^{}(1x^{})},`$ (13)
and the dynamical gluon mass $`\mu _{dyn}`$ is given by Eq. (9); also the following momenta are used $`Q^2=(Q_1^2+Q_2^2)/2`$ and $`\delta Q^2=(Q_1^2Q_2^2)/2`$. Calculating Eq. (10) with an explicit form of similarity function, gives in the leading order $`\delta Q^2Q^2`$ the following $`q\overline{q}`$-interaction $`^\mathrm{?}`$
$`V_{q\overline{q}}=\gamma ^\mu \gamma _\mu \left(C_f\alpha _s{\displaystyle \frac{4\pi }{Q^2}}+\sigma {\displaystyle \frac{8\pi }{Q^4}}\right),`$ (14)
where to this order $`Q^2`$ reduces in the instant frame to the square of gluon momentum transfer, $`Q^2\stackrel{}{q}^2`$, with the gluon momentum $`\stackrel{}{q}=(q_z,q_{})`$. In order to reproduce the standard Coulomb and linear confining potentials, the correct prefactors are restored, using the freedom to fit $`Const`$ and $`\sigma `$ terms. Confining term in Eq. (14), with singular behavior like $`1/\stackrel{}{q}^4`$, arise from the elimination of the quark-gluon coupling at small gluon momenta, that is governed by the cutoff dependent, dynamical gluon mass.
## Acknowledgments
The author would like to thank the organizers of the workshop for hospitality and support. The author is thanful to Stan Brodsky and Lev Lipatov for helpful discussions. This work was supported by DOE grants DE-FG02-96ER40944, DE-FG02-97ER41048 and DE-FG02-96ER40947.
## References |
warning/0003/astro-ph0003001.html | ar5iv | text | # The Baryonic Tully-Fisher Relation
## 1 Introduction
The relation between luminosity and rotation velocity for galaxies is well known (Tully & Fisher 1977). It has been used extensively in estimating extragalactic distances (e.g., Sakai et al. 2000, Tully & Pierce 2000), and it provides a critical constraint on galaxy formation theory (Dalcanton, Spergel, & Summers 1997; McGaugh & de Blok 1998; Mo, Mao, & White 1998; Steinmetz & Navarro 1999; van den Bosch 1999). However, the physical basis of the Tully-Fisher relation remains unclear.
The requirements of the empirical Tully-Fisher relation are simple, but the steep slope and small scatter are difficult to understand. Luminosity must trace total (dark plus luminous) mass, which in turn scales exactly with circular velocity. Considerable fine-tuning is required to obtain these strict proportionalities (McGaugh & de Blok 1998). The intrinsic properties of dark halos are not expected to be as tightly correlated as observed (Eisenstein & Loeb 1995). The mapping from the properties of dark matter halos to observable quantities should introduce more scatter, not less. Somehow the baryons โknowโ precisely how many stars to form.
Let us suppose that, for whatever fundamental reason, there does exist a universal relationship between total mass and rotation velocity of the form $`_{tot}V_c^b`$. The empirical Tully-Fisher relation then follows if luminosity traces mass:
$$L=\mathrm{{\rm Y}}_{}^1f_{}f_df_b_{tot},$$
(1)
where $`f_b`$ is the baryon fraction of the universe, $`f_d`$ is the fraction of the baryons associated with a particular galaxy halo which reside in the disk, $`f_{}`$ is the fraction of disk baryons in the form of stars, and $`\mathrm{{\rm Y}}_{}`$ is the mass-to-light ratio of the stars. Each of the pieces which intervene between $`L`$ and $`_{tot}`$ must be a nearly universal constant shared by all disks in order to maintain the strict proportionality the Tully-Fisher relation requires. Cast in this form, the traditional luminosity-linewidth relation is a sub-set of a more fundamental relation between baryonic mass and rotational velocity. In this context, one would expect to find galaxies which deviate from the luminosity-linewidth relation because much of their baryonic mass is not in the form of stars. For example, a gas rich galaxy should appear underluminous for its circular velocity, but would, after correction for the gas content, fall on the underlying โBaryonic Tully-Fisher relationโ (cf. Freeman 1999).
In this paper, we specifically test this premise by constructing the luminosity-linewidth and Baryonic Tully-Fisher relations for a sample of late type galaxies that span a much larger range of luminosities than any previously available sample. Section 2 describes the data we employ. Section 3 discusses the results and ยง4 explores some of their implications. A summary is given in ยง5. All distance dependent quantities assume $`H_0=75\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$.
## 2 Data
We employ several data sets to maximize the dynamic range over which we can explore the Tully-Fisher relation. The different data sets have photometry in different pass bands. To put the data on the same system, and get at the question of the underlying mass, we assume a stellar mass-to-light ratio for each pass band. Stellar mass is most directly traced by the redder pass bands, so we adopt these when possible.
For galaxies with $`V_c100\mathrm{km}\mathrm{s}^1`$ we use the extensive $`H`$-band data for late type cluster spirals of Bothun et al. (1985). Circular velocities are estimated as half of the linewidth $`W_{20}`$. For galaxies of lower rotation velocity, we use the data for late type dwarf low surface brightness galaxies from the survey of Schombert, Pildis, & Eder (1997). This is currently the largest sample of field dwarf galaxies with both linewidths $`W_{20}`$ and HI masses (Eder & Schombert 1999) and red band photometry (Pildis, Schombert, & Eder 1997). The photometry provides $`I`$-band magnitudes and axial ratios for inclination estimates. Of these galaxies, only those with axial ratios $`b/a<0.71`$, corresponding to $`i>45^\mathrm{ยฐ}`$ for an intrinsic axial ratio of $`q_0=0.15`$, are used in order to minimize $`\mathrm{sin}(i)`$ errors. These nevertheless contributes substantially to the scatter, as inclinations estimated from the axial ratios of dim galaxies are intrinsically uncertain. The data for these faint galaxies extend the Tully-Fisher relation to much lower luminosities and circular velocities than have been explored previously.
The fundamental rotation velocity of interest here is the flat portion of the rotation curve, $`V_{flat}`$. Presumably, the line width $`W_{20}`$ commonly employed in Tully-Fisher work is an adequate indicator of $`V_{flat}`$. As a check on this, we also employ the data of Verheijen (1997) and McGaugh & de Blok (1998) for which $`V_{flat}`$ is measured from resolved rotation curves. The data of Verheijen (1997) are $`K^{}`$-band data for spiral galaxies in the UMa cluster (Tully et al. 1996), while the data discussed by McGaugh & de Blok (1998) are $`B`$-band data drawn from a variety of sources.
The two red-band data sets, the $`H`$-band data of Bothun et al. (1985) at the bright end, and the $`I`$-band dwarf galaxy sample at the faint end, together suffice to define a Tully-Fisher relation over five decades in stellar mass. The rotation curve samples are consistent with these data. For comparison, we also examine the gas rich, late type galaxy sample of Matthews, van Driel, & Gallagher (1998). Their $`B`$-band data are entirely consistent with our own data, provided we make the same inclination cut, $`i>45^\mathrm{ยฐ}`$. Though this inclination limit is of obvious importance, it is interesting to note that including or excluding the galaxies they note as having strongly asymmetric or single-horned HI profiles makes no difference to the result.
In all cases, we have simply taken the data as given by each source. Aside from the necessary inclination correction, we have not made any corrections for internal extinction or for non-circular motions (shown to be small for late type systems by Rix & Zaritsky 1995 and by Beauvais & Bothun 1999). That the data treated in this way produce a good Tully-Fisher relation indicates, to first order, that these effects are not important.
## 3 Results
Figure 1 illustrates the Tully-Fisher relation for the combined data sets. Two versions are shown: in (a) the stellar mass is plotted in place of luminosity, and in (b) the total luminous baryonic mass is shown. In order to place the data sets using different band passes for photometry on the same scale, we convert luminosity to stellar mass assuming a fixed mass-to-light ratio $`\mathrm{{\rm Y}}_{}`$ for each band. The value of the mass-to-light ratio appropriate to the stellar populations of late type galaxies with ongoing star formation has been examined in detail by de Jong (1996). We adopt his model for a 12 Gyr old, solar metallicity population with a constant star formation rate and Salpeter IMF. The adopted mass-to-light ratios are $`\mathrm{{\rm Y}}_{}^B=1.4`$, $`\mathrm{{\rm Y}}_{}^I=1.7`$, $`\mathrm{{\rm Y}}_{}^H=1.0`$, and<sup>1</sup><sup>1</sup>1For the mean $`HK^{}`$ color of late type galaxies given by de Jong (1996), $`\mathrm{{\rm Y}}_{}^H=1.2\mathrm{{\rm Y}}_{}^K^{}`$. $`\mathrm{{\rm Y}}_{}^K^{}=0.8_{\mathrm{}}/L_{\mathrm{}}`$. These $`K^{}`$ and $`I`$-band mass-to-light ratios are consistent with the maximum disk fits to the bright galaxies of Verheijen (1997) and Palunas (1996). We do of course expect variation in stellar populations and their mass-to-light ratios. This should be modest in the redder bands, especially $`H`$ and $`K^{}`$, which are not very sensitive to differences in star formation history. The $`I`$-band mass-to-light ratio is not very sensitive to metallicity (Worthey 1994), so this should suffice for the fainter galaxies which in any case are dominated by gas mass. The $`B`$-band is a less robust indicator of stellar mass, so we do not include these data in the fit in Figure 1(b). While the absolute normalization of stellar mass-to-light ratios remains uncertain, tweaking the adopted values has no effect on the basic result.
The stellar mass plotted in Figure 1(a) is simply $`_{}=\mathrm{{\rm Y}}_{}L`$, so this plot is directly analogous to the conventional luminosity-linewidth diagram. The baryonic disk mass plotted in Figure 1(b) is the sum of stars and gas, $`_d=_{}+_{gas}`$. The mass in gas is taken from the observed HI mass with the standard correction for helium and metals: $`_{gas}=1.4_{HI}`$. It appears that molecular gas is not a significant mass component in these late type galaxies (Schombert et al. 1990; de Blok & van der Hulst 1998; Mihos, Spaans, & McGaugh 1999; Gerritsen & de Blok 1999).
There have long been hints (e.g., Romanishin, Strom, & Strom 1983) that faint galaxies fall below the extrapolated Tully-Fisher relation for bright galaxies. Matthews et al. (1998) and Stil & Israel (1999) claim to see this in their samples. However, it is not clear from their data in Figure 1. The apparent discrepancy in our results stems not from a difference in the data, but from what is taken to define the Tully-Fisher relation. Matthews et al. (1998) and Stil & Israel (1999) compare their data to lines fit to the $`B`$-band data of brighter galaxies. These fiducial lines have a shallow slope which considerably overpredicts the luminosities of faint galaxies when extrapolated to low circular velocity. It is not clear that it is safe to extrapolate the slope in this fashion. Extinction appears to be relatively more important in brighter galaxies, with careful corrections giving steeper slopes (Tully et al. 1998). Samples of galaxies with low intrinsic extinctions also give considerably steeper $`B`$-band slopes (Sprayberry et al. 1995; Verheijen 1997; McGaugh & de Blok 1998). The $`H`$-band data of Bothun et al. (1985) and the $`K^{}`$-band data of Verheijen (1997), two bands where extinction is minimal, also indicate steep slopes. A steep slope is also supported by the calibration of the Tully-Fisher relation from the HST Key Project (Sakai et al. 2000). Such a slope eliminates the discrepancy reported by Matthews et al. (1998) and by Stil & Israel (1999).
Nevertheless, it is now clear from the larger dwarf sample employed here that there is indeed a break in the Tully-Fisher relation for faint field<sup>2</sup><sup>2</sup>2The $`K^{}`$-band data of Pierini & Tuffs (1999) shows a steep slope with no break down to $`V_c60\mathrm{km}\mathrm{s}^1`$. These are cluster galaxies, so this makes sense if these objects are less gas rich than the field sample. galaxies. For $`V_c90\mathrm{km}\mathrm{s}^1`$, galaxies are underluminous for their rotation velocity as predicted by the extrapolation of a linear fit to the bright galaxy data. There is a great amount of scatter here too โ the relation bends and flares. There have been concerns that there might be curvature in the Tully-Fisher relation (e.g., Bothun & Mould 1987) but the data in Figure 1(a) are probably better described by a broken power law, if it makes sense to fit anything to the faint end at all.
A break in the Tully-Fisher relation would have important ramifications for its application and interpretation. However, many of these faint galaxies are very gas rich. So much so, in fact, that the gas outweighs the stars in most of them for any reasonable choice of stellar mass-to-light ratio (Schombert, McGaugh, & Eder 2000). Therefore, we examine in Figure 1(b) the effects of including the gas mass in the ordinate by plotting the total observed baryonic disk mass, $`_d=_{}+_{gas}`$. This has the remarkable effect of restoring a single linear relation over the entire span of the observations.
It appears that the fundamental relation underpinning the Tully-Fisher relation is one between rotation velocity and total baryonic disk mass. This relation has the form
$$_d=๐V_c^b.$$
(2)
An unweighted fit to the red ($`I`$, $`H`$, and $`K^{}`$-band) data gives $`\mathrm{log}๐=1.57\pm 0.25`$ and $`b=3.98\pm 0.12`$. The precise value of the normalization would of course change if we assumed a different distance scale or different stellar mass-to-light ratios. The slope is indistinguishable from $`b=4`$. If we fix the slope to this value, the normalization is $`๐35(\mathrm{{\rm Y}}_{}^K^{}/0.8)h_{75}^2_{\mathrm{}}\mathrm{km}^4\mathrm{s}^4`$.
## 4 Implications
The basic result seen in Figure 1(b) falls directly out of the observations. All we have done is assume a plausible mass-to-light ratio for the stars, added in the gas mass, and plotted the data. This simple result has a number of interesting implications.
First, there is an apparently universal relation between baryonic mass and rotation velocity, with a single normalization. While this relation specifically applies to our sample of late type spiral galaxies, it seems plausible that it might also apply to early type spirals, provided appropriate consideration is given to the bulge component, which might require a different $`\mathrm{{\rm Y}}_{}`$, and to any other baryonic components which might be significant (like molecular gas).
The logarithmic slope of the relation is indistinguishable from 4. While this slope is often attributed to the virial theorem, it is possible to derive other slopes as well depending on the assumptions one makes (Mo et al. 1998). Current cold dark matter models predict a slope of 3 (Mo et al. 1998; Steinmetz & Navarro 1999) which is excluded at $`8\sigma `$. Significant tweaking is required to obtain the observed slope. Feedback from supernovae is often invoked in this context (van den Bosch 1999), but it is not obvious that the modest amount of feedback required by the Tully-Fisher relation is consistent with the large amount needed to explain the luminosity function (Lobo & Guiderdoni 1999). The correct slope and normalization is predicted by one alternative to cold dark matter (Milgrom 1983). In this alternative there is no dark matter โ all of the mass is baryonic.
Whatever mechanism sets the observed relation is intimately connected to the observed baryonic mass. The interpretation of the standard luminosity-linewidth relation has long supposed that the stellar mass-to-light ratios of galaxies are a nearly uniform. Indeed, the error budget allowed by the modest amount of intrinsic scatter observed in the $`K^{}`$-band is easily consumed by variations in the star formation history (Verheijen 1997). There is little room left for variation in the IMF, or cosmic scatter in the underlying mass-rotation velocity relation.
We have now addressed another piece of this puzzle. In addition to the near constancy of $`\mathrm{{\rm Y}}_{}`$, we have explicitly corrected for the stellar fraction $`f_{}`$. Equation (1) now reduces to
$$_d=f_df_b_{tot}.$$
(3)
The presumed mass-rotation velocity relation can now show through in the observations provided both $`f_b`$ and $`f_d`$ are universal constants. The baryon fraction of the universe is constant by definition. But it is less obvious that the fraction of baryons which reside in the disk should be the same for all spirals. Indeed, it is frequently suggested (e.g., Navaroro, Eke, & Frenk 1996) that the sort of faint dwarfs studied here are likely to lose a significant portion of their baryons. This idea is blatantly at odds with the data, as the product $`f_df_b`$ would no longer<sup>3</sup><sup>3</sup>3One could contemplate a variable $`f_d`$ provided that it was a very finely tuned (zero scatter) function of circular velocity. For example, $`f_dV_c`$ would recover the slope predicted by CDM. be constant.
It seems to us implausible that $`f_d`$ could be some arbitrary yet universal fraction. While it is easy to imagine mechanisms which might prevent some of the baryons from cooling to join the disk, it is difficult to contemplate any which do so with the required precision. There is very little room in the budget for the intrinsic scatter for any scatter in $`f_d`$. Let us call the mass in baryons not already accounted for in the disk mass $`_{other}`$. The disk fraction is then
$$f_d=\frac{_{}+_{gas}}{_{}+_{gas}+_{other}}.$$
(4)
If this other form of baryonic mass is significant ($`_{other}_{}`$), then $`f_d<1`$, but there should be a lot of scatter in $`f_d`$ unless some magical mechanism strictly regulates the ratio $`_{other}/(_{}+_{gas})`$. This unlikely situation occurs naturally only if $`_{other}_{}+_{gas}`$, so $`f_d1`$. The modest intrinsic scatter in the Baryonic Tully-Fisher relation therefore suggests that the luminous mass in stars and gas represents nearly all the baryons associated with an individual galaxy and its halo, arguing against a significant mass of dark baryons in these systems.
## 5 Conclusions
We have explored the Tully-Fisher relation over five decades in luminous mass. This is a considerable increase in dynamic range over previous studies. We find clear evidence for a break in the optical Tully-Fisher relation around $`V_c90\mathrm{km}\mathrm{s}^1`$. Galaxies with rotation velocities less than this are underluminous relative to the extrapolation of the fit to more rapidly spinning galaxies. However, these faint galaxies are very gas rich. Considering both stellar and gas mass restores a linear relation over the entire observed range.
These observations strongly suggest that the Tully-Fisher relation is fundamentally a relation between rotation velocity and total baryonic disk mass. This relation has the form
$$_d=๐V_c^4$$
with $`๐35(\mathrm{{\rm Y}}_{}^K^{}/0.8)h_{75}^2_{\mathrm{}}\mathrm{km}^4\mathrm{s}^4`$. The well known optical Tully-Fisher relation is an approximation to this more fundamental relation in the limit of galaxies dominated by stars.
The existence of the Baryonic Tully-Fisher relation has a number of important implications. That it works means that stars in spiral galaxies have mass-to-light ratios which are reasonable for composite stellar populations. The modest amount of scatter indicates that the IMF must be nearly universal in order to yield such uniform mass-to-light ratios. Only corrections for gas content are necessary to obtain the Baryonic Tully-Fisher relation. The data do not allow much room for any further significant baryonic mass components. Any component of dark baryons which does exist must do so in strict proportionality to the observed baryons, with effectively zero scatter. This unlikely situation argues against a significant mass in dark baryons in any form (be it very cold molecular gas in the disk, very hot ionized gas in the halo, or baryonic MACHOs). Any model which supposes a large mass of such baryons must explain why it contributes so little to the scatter in the Baryonic Tully-Fisher relation.
The results presented here make sense in terms of a simple interpretation of the Tully-Fisher relation in which the mass of observed baryons is directly proportional to the total mass which in turn scales with the observed rotation velocity. This potentially includes the case where the mass observed in baryons is the total mass (Milgrom 1983). Matching these observations is a substantial challenge for modern structure formation theories based on cold dark matter. These predict a slope which is too shallow (3 rather than 4, different by $`8\sigma `$), and fail to anticipate that effectively all the baryons associated with a halo have cooled into the disk.
We thank Ken Freeman for relevant conversations, and the referee, H.-W. Rix, for many insightful comments. The work of SSM is supported in part by NSF grant AST 99-01663. ADDENDUM:
The text above is rigorously identical to that accepted by ApJ Letters. Two points were lost in the effort to meet the Lettersโ page limit.
1: Maximum Disk Mass-to-Light Ratios: We have assumed a constant $`\mathrm{{\rm Y}}_{}`$ which is plausible for a composite stellar population. Presumably, there is scatter about this value which is reflected in the non-observational component of the scatter in Figure 1(b). If instead we adopted the stellar mass-to-light ratio suggested by maximum disk fits to rotation curves, the resultant relation would have a much larger scatter. The reason for this is that $`\mathrm{{\rm Y}}_{}^{max}`$ increases systematically with decreasing disk central surface brightness (de Blok & McGaugh 1997, MNRAS, 290, 533; Swaters, Madore, & Trewhella 2000, ApJ, 531, L107), yet there is a range of surface brightnesses at a given luminosity. For example, Swaters et al. discuss a number of galaxies with $`M_B18`$ for which $`\mathrm{{\rm Y}}_{}^{max}`$ ranges from 1.5 to 17. Obviously, this factor would cause a large difference in the computed mass at a given circular velocity, and inflate the scatter in the Baryonic Tully-Fisher Relation. This argues against maximal disks in low surface brightness galaxies, though these remain plausible for high surface brightness galaxies.
2: Deviant Galaxies: Recently, OโNeil, Bothun, & Schombert (1999, AJ, 119, 136) have pointed out a population of red low surface brightness galaxies. Four of these galaxies deviate significantly from the Baryonic Tully-Fisher relation, falling roughly an order of magnitude below it (having too little mass for their circular velocity). Given their red colors, these galaxies might have high $`\mathrm{{\rm Y}}_{}`$, or could contain some other component of baryonic mass. This point is tentative, so here we note only that the Baryonic Tully-Fisher relation is a good diagnostic for identifying unusual galaxies. This also applies to galaxies which might have very low $`\mathrm{{\rm Y}}_{}`$ as a result of a young stellar population (e.g., Meurer et al. 1996, AJ, 111, 1551). |
warning/0003/hep-th0003206.html | ar5iv | text | # Inflationary Universe in Higher Derivative Induced Gravity
## Abstract
In an induced-gravity model, the stability condition of an inflationary slow-rollover solution is shown to be $`\varphi _0_{\varphi _0}V(\varphi _0)=4V(\varphi _0)`$. The presence of higher derivative terms will, however, act against the stability of this expanding solution unless further constraints on the field parameters are imposed. We find that these models will acquire a non-vanishing cosmological constant at the end of inflation. Some models are analyzed for their implication to the early universe.
PACS numbers: 98.80.Cq; 04.20 -q;
In a scale-invariant model, all dimensionful parameters are functionals of the scalar field. Therefore, scale invariance provides a natural way resolving the physical origin of these dimensionful parameters.
Scale invariance is also known to be important in various branches of physics. For example, QCD and many other inflationary models have been studied in the literature . Note that inflation resolves many problems of the standard big bang cosmology . These problems include the flatness, the monopole, and the horizon problem. In addition, local scale (or Weyl) symmetry has been suggested to be related to the missing Higgs problem in electro-weak theory . Weyl symmetry has also been the focus of many recent activities . Scale-invariant effective theory is also suggested to be important for the physics near fixed points of the renormalization group trajectory .
In addition, higher derivative terms should be important for the physics near the Planck scale . For example, higher-order corrections derived from the quantum gravity or the string theory have been considered in the study of the inflationary universe . Higher derivative terms also arise as quantum corrections to the matter fields . Moreover, the stability analysis of the pure higher-derivative models was shown in Ref. . It is hence interesting to extend this stability analysis to different models. In an induced-gravity model, it turns out that stability conditions of an inflationary solution are that the scalar field must obey a set of scale-invariant conditions under the slow-rollover approximation. We will also study the implication of this constraint to the inflationary universe in this paper.
We will focus on the induced-gravity model with an $`R^2`$ coupling given by
$$S=d^4x\sqrt{g}\left\{\frac{1}{2}ฯต\varphi ^2R\frac{1}{2}g^{\mu \nu }_\mu \varphi _\nu \varphi V(\varphi )\frac{\alpha }{3}R^2\right\}$$
(1)
in this paper. Here $`ฯต`$ and $`\alpha `$ are dimensionless coupling constants. $`V(\varphi )`$ is any possible symmetry-breaking potential. Note that there are additional fourth-derivative terms in the most general higher derivative theory. They are related to the $`R^2`$ term, due to the Euler constraint and the fact that the Weyl tensor vanishes in the Friedmann-Robertson-Walker (FRW) spaces , in four-dimensional spaces.
We will define $`sV_04V_0\varphi _0_\varphi V(\varphi _0)`$ with $`s`$ denoting the scaling factor, $`\varphi _0`$ denoting the initial condition of the inflaton $`\varphi `$, and $`V_0V(\varphi _0)`$. In addition, the case that $`s=0`$ will be referred to as the scaling condition in this paper. Note that the scaling condition will be shown to be a direct consequence of the slow-rollover approximation. Hence the initial data has to be close to the scaling condition. We are going to show, however, that the induced-gravity model tends to stabilize the inflationary phase. This is true if initial conditions of the inflaton are close to the scaling condition. The presence of the higher derivative term (HDT) will further impose strong constraints on field parameters and scalar potentials. These constraints are required to generate a stable inflationary phase. Otherwise, this theory can not permit an exponentially expanding solution under the scaling condition in the presence of the HDT.
Note that our universe is homogeneous and isotropic to a very high degree of precision . Such a universe is described by the well-known FRW metric . Therefore, we will work on the FRW metric that can be read off directly from the following equation:
$$ds^2g_{\mu \nu }dx^\mu dx^\nu =dt^2+a^2(t)\left(\frac{dr^2}{1kr^2}+r^2d\mathrm{\Omega }\right).$$
(2)
Here $`d\mathrm{\Omega }`$ is the solid angle $`d\mathrm{\Omega }=d\theta ^2+\mathrm{sin}^2\theta d\chi ^2`$ and $`k=\mathrm{\hspace{0.17em}0},\pm 1`$ stand for a flat, closed, and open universe respectively.
The Friedmann equation can be shown to be
$`3ฯต\varphi ^2(H^2+{\displaystyle \frac{k}{a^2}}+2H{\displaystyle \frac{\dot{\varphi }}{\varphi }})=V+{\displaystyle \frac{1}{2}}\dot{\varphi }^2+K.`$ (3)
Here $`K12\alpha [2H\ddot{H}+6H^2\dot{H}\dot{H}^22H^2k/a^2+k^2/a^4]`$ denotes the contribution from the HDT. Moreover, the Euler-Lagrange equation for $`\varphi `$ is
$$\ddot{\varphi }+3H\dot{\varphi }+\frac{V}{\varphi }=6ฯต\varphi (\dot{H}+2H^2+\frac{k}{a^2}).$$
(4)
Note that the HDT does not affect the $`\varphi `$ equation directly as shown above.
One will analyze the inflationary solution under the slow-rollover approximation such that $`|\dot{\varphi }/\varphi |H`$, and $`|\ddot{\varphi }/\varphi |H^2`$ for a brief period of time. The slow-rollover approximation will also be shown to be consistent with field equations. Assuming that $`\varphi =\varphi _0`$ and $`H=H_0+\delta H`$, one can perturb the Friedmann equation and the scalar field equation. These perturbed equations can be employed to study the stability of the inflationary solution. Accordingly, one can show that leading-order perturbation equations give
$`3ฯต\varphi _0^2H_0^2`$ $`=`$ $`V_0,`$ (5)
$`\varphi _0{\displaystyle \frac{V}{\varphi }}(\varphi =\varphi _0)`$ $`=`$ $`12ฯต\varphi _0^2H_0^2.`$ (6)
Therefore, the initial data satisfies the scaling condition in this approximation. In addition, linear-order perturbation equations give
$`4\alpha \delta \ddot{H}+12\alpha H_0\delta \dot{H}ฯต\varphi _0^2\delta H`$ $`=`$ $`0,`$ (7)
$`\delta \dot{H}+4H_0\delta H`$ $`=`$ $`0.`$ (8)
Therefore, one has
$`\delta H`$ $``$ $`\mathrm{exp}(4H_0t),`$ (9)
$`ฯต\varphi _0^2`$ $`=`$ $`16\alpha H_0^2.`$ (10)
Eq. (7) will not be present without the HDT. Therefore, one will not have the constraint (10) accordingly. Moreover, Eq. (9) indicates that inflation tends to stabilize the inflationary phase under the scaling condition. Note that Eq. (10) is the extra constraint derived from the HDT. This indicates that the gravitational constant ($`ฯต\varphi _0^2/2`$) is related to the Hubble constant $`H_0`$ during the inflationary phase. Therefore, a physically acceptable inflationary induced-gravity model will be affected significantly by the HDT.
In addition, the first-order perturbation equation shows that the inflationary solution is indeed stable against the perturbation $`\delta H`$. Therefore, inflation will remain effective for at least a brief moment while $`\varphi `$ changes slowly. Note also that the $`\varphi `$ equation states that
$$\ddot{\varphi }+3H_0\dot{\varphi }0$$
during the period when $`HH_0`$. This gives
$$\varphi \varphi _0+\frac{\dot{\varphi }_0}{3H_0}[1\mathrm{exp}(3H_0t)].$$
(11)
Therefore, the slow-rollover approximation is indeed consistent with field equations. Consequently, if the initial data satisfies the scaling condition, the system will undergo a strong inflationary process and remain stable for a long period of time under the scaling condition. Therefore, we will focus on the case that the initial data of the effective theory obeys the $`s=0`$ condition.
Note that leading-order perturbation equations give us a few constraints on the field parameters according to
$$4V_0=\varphi _0\frac{V}{\varphi }(\varphi =\varphi _0)=12ฯต\varphi _0^2H_0^2=192\alpha H_0^4.$$
(12)
This is equivalent to
$`H_0^2`$ $`=`$ $`{\displaystyle \frac{ฯต\varphi _0^2}{16\alpha }},`$ (13)
$`4V_0`$ $`=`$ $`\varphi _0{\displaystyle \frac{V}{\varphi }}(\varphi =\varphi _0)={\displaystyle \frac{3ฯต^2}{16\alpha }}\varphi _0^4.`$ (14)
Therefore, there are indeed strong constraints on the possible form of the scalar field potential according to Eq. (14). These constraints also relate the field parameters in a nontrivial way. We will come back to this point later and study the constraint equation for some extended $`\varphi ^4`$ models.
If the scaling condition is not obeyed closely, the inflationary solution will not be strictly stable. This will act in favor of the graceful-exit process. In such cases, the scalar field will obey the following equation
$`\ddot{\varphi }`$ $`+`$ $`3H\dot{\varphi }+{\displaystyle \frac{\dot{\varphi }^2}{\varphi }}+\left[_\varphi V4V/\varphi \right]/(1+6ฯต)`$ (15)
$`=`$ $`\left[\dot{K}+4HK\right]/(1+6ฯต)\varphi `$ (16)
which can be derived from differentiating Eq. (3) and comparing it with Eq. (4). Note further that Eq. (16) is equivalent to the $`G_{ij}`$ component of the Einstein equation. Even this equation is redundant, it is still very useful for our analysis. In summary, the inflationary solution can not be stable unless (i) the scaling condition is closely obeyed and (ii) $`\alpha `$ is constrained by Eq. (10). In such cases, the dynamics of the scalar field can be depicted from Eq. (16). We are about to show that the case (ii) will be violated in the conventional $`\varphi ^4`$ SSB potential. Therefore, the system will follow the evolutionary process similar to the one described in Ref. . On the other hand, the physics will be different when the initial data falls too close to the scaling condition.
For comparison, the equation of motions will become
$`(H^2+{\displaystyle \frac{k}{a^2}}+2H{\displaystyle \frac{\dot{\varphi }}{\varphi }})=V+{\displaystyle \frac{1}{2}}\dot{\varphi }^2+K,`$ (17)
$`\ddot{\varphi }+3H\dot{\varphi }+{\displaystyle \frac{V}{\varphi }}=0`$ (18)
for the $`R^2`$-corrected Einstein theory given by $`=R/2\alpha R^2+_\varphi `$. Hence the scaling constraint no longer holds here since the $`\varphi `$-equation does not couple to the $`R^2`$ term directly. Indeed, one can show that $`H_0^2=V_0/3`$ from the zeroth-order perturbation equation. Note that the perturbation is done with respect to $`H=H_0+\delta H`$ under the slow-rollover approximation $`|\ddot{\varphi }|H|\dot{\varphi }|`$ and $`H|\dot{\varphi }|`$. One can also show that the first-order perturbation equation gives Eq. (8) after setting $`ฯต\varphi _0^21`$. Therefore, the effect of the $`R^2`$-corrected Einstein theory is different from the induced-gravity model. Hence the scale-invariant initial condition is a very unique property of induced-gravity models.
For a physical application, we will consider the following effective symmetry-breaking potential
$$V=\frac{\lambda _1}{4}(\varphi ^2v^2)^2+\frac{\lambda _2}{4}\varphi ^4\mathrm{\Lambda }.$$
(19)
We are about to show that the apparent cosmological-constant term $`\mathrm{\Lambda }`$ has to be non-vanishing in order to admit a consistent inflationary solution. In addition, it reduces to the standard $`\varphi ^4`$ SSB potential if $`\lambda _2=\mathrm{\Lambda }=0`$. Therefore, neither scale-invariant potential nor standard $`\varphi ^4`$ SSB potential can provide a physically acceptable inflationary solution under the influence of the HDT. Hence one has to introduce an alternative asymmetric potential in order to generate an inflationary solution.
Indeed, one can solve Eq. (14) and show that
$$\varphi _0^2=v^2\frac{4\mathrm{\Lambda }}{\lambda _1v^2}=\frac{16\alpha \lambda _1v^2}{16\alpha (\lambda _1+\lambda _2)3ฯต^2}.$$
(20)
Therefore, one can derive
$$\mathrm{\Lambda }=\frac{\lambda _1}{4}v^4\left[\frac{16\alpha \lambda _23ฯต^2}{16\alpha (\lambda _1+\lambda _2)3ฯต^2}\right].$$
(21)
Writing $`\lambda \lambda _1+\lambda _2`$, one can further show that the extended $`\varphi ^4`$ SSB potential reads
$$V=\frac{\lambda }{4}\varphi ^4(\frac{\lambda }{2}\frac{3ฯต^2}{32\alpha })\varphi _0^2\varphi ^2+(\frac{\lambda }{4}\frac{3ฯต^2}{64\alpha })\varphi _0^4.$$
(22)
This is the form of the most general extended $`\varphi ^4`$ SSB potential that could admit an inflationary solution. One can also show that the minimum of this potential is $`V_mV(\varphi _m)=(3ฯต^2/16\alpha )[1/43ฯต^2/64\alpha \lambda ]\varphi _0^4`$ when $`\varphi ^2=\varphi _m^2(13ฯต^2/16\alpha \lambda )\varphi _0^2`$. In addition, one has $`V_m=(3ฯต^2/16\alpha \lambda )V(0)`$ where $`V(0)V(\varphi =0)`$ is the maximum of $`V`$. Note also that $`V_m<V(0)`$ is consistent with the equation $`\varphi ^2=\varphi _m^2(13ฯต^2/16\alpha \lambda )\varphi _0^2`$. This implies that $`3ฯต^2/16\alpha \lambda <1`$. Hence one has $`\alpha >0`$ because that $`V_0=3ฯต\varphi _0^2H_0^2=3ฯต^2\varphi _0^4/64\alpha >0`$. In addition, one expects $`\lambda >0`$ since $`V^{\prime \prime }(\varphi _m)=2\lambda \varphi _m^2>0`$ for a local minimum at $`\varphi _m`$. Therefore, one shows that $`V_m=3ฯต^2\varphi _0^2\varphi _m^2/64\alpha >0`$.
Moreover, the effective gravitational constant observed in the post-inflationary phase is related to $`\varphi _0`$ by the identity $`1/4\pi G=ฯต\varphi _m^2=(13ฯต^2/16\alpha \lambda )ฯต\varphi _0^2`$. In addition, the effective cosmological constant observed in the post-inflationary phase is $`V_m=3H_0^2/2`$. Here we have set $`ฯต\varphi _m^2/2=1`$ in Planck unit. If the scale factor $`a(t)`$ is capable of expanding some $`60`$ e-fold in a time interval of roughly $`\mathrm{\Delta }T10^8`$ Planck unit, the Hubble constant should be of the order $`H_0^210^6`$ in Planck unit. Therefore, one ends up with a rather big cosmological constant of the order $`10^6`$ if the extended $`\varphi ^4`$ model is in effect.
One can now show that the case $`\mathrm{\Lambda }=\lambda _2=0`$ is problematic. Indeed, $`\mathrm{\Lambda }=0`$ implies that $`\lambda _2=3ฯต^2/16\alpha `$ from Eq. (21). Hence $`ฯต/\alpha =0`$ if $`\lambda _2=0`$. This is apparently inconsistent with our assumption that $`\alpha `$ is small and $`ฯต`$ is finite. In fact, the case that $`ฯต=0`$ will lead to an infinite gravitational constant. Hence it should be ruled out. Therefore, the case that $`\mathrm{\Lambda }=\lambda _2=0`$ can not support an inflationary phase if the HDT is present.
Note that the expansion rate $`H_0=\sqrt{ฯต\varphi _0^2/16\alpha }`$ can be adjusted to accommodate $`60`$ e-fold expansion rather easily . Note also that small $`\alpha `$, hence small higher-order correction, will act in favor of the inflationary process. In addition, the slow-rollover approximation is taken care of automatically by the higher-order term. Therefore, the only constraint on $`\varphi _0`$ is that it should be small according to Eq. (11).
Once the scalar field rolls down toward the minimum-potential state, inflation will come to an end. In addition, the soft-expansion era in the post-inflationary phase will be dominated by another lower-order induced-gravity model . Therefore, the re-heating process will be taken over by that lower-order effective induced-gravity model . Hence the HDT, acting in favor of the inflation process, plays an important role during the inflationary phase. It is still true even if the higher-order correction is small, namely, $`\alpha 1`$. Note, however, that this model implies a non-vanishing cosmological constant. This may have to do with the field contents of the early universe .
In other words, the smallness of the cosmological constant is not resolved by this approach. Something else has to help resolving the cosmological-constant problem. One possibility already mentioned earlier is that this induced-gravity theory remains effective, only during the inflationary era, as an collective effect of the physics in the early universe . This effective induced theory will no longer be held responsible for the physics after inflation is completed.
In addition, one can also consider the following symmetry-breaking Coleman-Weinberg potential from radiative correction
$$V=\frac{\lambda _1}{4}\varphi ^4\mathrm{ln}(\frac{\varphi }{v})^4+\frac{\lambda _2}{4}\varphi ^4\mathrm{\Lambda }.$$
(23)
Note that we will use the same notation for $`V`$, $`\lambda _i`$, $`\mathrm{\Lambda }`$, etc. for simplicity although we are working on a different model. One can show that the first constraint in Eq. (14) gives $`\mathrm{\Lambda }=\lambda _1\varphi _0^4/4`$. The second one gives $`V_0=3ฯต^2\varphi _0^4/64\alpha `$. This implies that $`\varphi _0=v\mathrm{exp}[3ฯต^2/64\alpha \lambda _1(\lambda _1+\lambda _2)/4\lambda _1]`$. Therefore, one can put the potential as
$$V=\frac{\lambda _1}{4}\varphi ^4\mathrm{ln}(\frac{\varphi }{\varphi _0})^4+\frac{3ฯต^2}{64\alpha }\varphi ^4\frac{\lambda _1}{4}(\varphi ^4\varphi _0^4).$$
(24)
In addition, one can show that the minimum state occurs when $`\varphi =\varphi _m=\varphi _0\mathrm{exp}[3ฯต^2/64\alpha \lambda _1]`$. Furthermore, one can show that $`V_m=(\lambda _1\varphi _0^4/4)\{1\mathrm{exp}[3ฯต^2/16\alpha \lambda _1]\}`$. Therefore, inflation can be achieved rather easily. On the other hand, one can show that $`V_m`$ depends on the choice of the parameters $`xฯต^2/\lambda _1`$ and $`y3/16\alpha `$ according to
$$V_m=\frac{e^{xy}1}{x}.$$
(25)
One can hence show that $`V_m`$ increases as $`x`$ decreases or $`y`$ increases. This is proved by showing that $`_xV_m`$ is always negative and $`_yV_m`$ is always positive definite. Hence by choosing smaller $`\lambda _1`$, larger $`\alpha `$, or larger $`ฯต`$ would lead to a smaller $`V_m`$. In practice, one should choose $`xy1`$ such that $`V_m3/16\alpha `$. In addition, the condition $`xy1`$ is equivalent to the condition $`\lambda _13ฯต^2/16\alpha =ฯต^2V_m(xy1)`$. Therefore, $`\alpha `$ has to be very large in order to push $`V_m`$ toward $`0`$. This is somewhat inappropriate as $`\alpha `$, related to the particle contents during inflation, can be computed from their quantum corrections in curved space . Therefore, one expects $`\alpha `$ to be small. Hence it is not likely that one can tune field parameters in order to push $`V_m`$ to the limit of observation in this theory. Note further that one has $`\alpha >0`$ because $`V_0=3ฯต\varphi _0^2H_0^2=3ฯต^2\varphi _0^4/64\alpha >0`$. In addition, one expects $`\lambda >0`$ since $`V^{\prime \prime }(\varphi _m)=4\lambda \varphi _m^2>0`$ where $`\varphi _m`$ is a local minimum. Hence one has $`V_m>0`$ if $`\alpha \lambda _1>0`$.
Note that the scaling condition is derived from the slow-rollover approximation. It was shown that the initial data has to be close to the scaling condition. We have, however, shown that a stable inflationary solution exists only when (i) scaling condition is closely obeyed and (ii) $`\alpha `$ is constrained by Eq. (10). We also show explicitly that two different extended models are not able to produce a universe with a vanishing cosmological constant all alone. Therefore, in the traditional $`\varphi ^4`$ SSB model under the scaling condition, inflationary solution does not favor a stable inflationary solution in the presence of the HDT. Accordingly, the traditional slow-rollover inflationary solution will soon fall off the scaling limit even if it started out close to the scaling condition. Hence, one does not need to worry about whether the scalar field will be frozen to the scaling condition. Therefore, the effect of the HDT will act in favor of the graceful-exit process.
Acknowledgments : This work is supported in part by the National Science Council under the contract number NSC88-2112-M009-001. |
warning/0003/cond-mat0003145.html | ar5iv | text | # Signatures of the excitonic memory effects in four-wave mixing processes in cavity polaritons
## Abstract
We report the signatures of the exciton correlation effects with finite memory time in frequency domain degenerate four-wave mixing (DFWM) in semiconductor microcavity. By utilizing the polarization selection rules, we discriminate instantaneous, mean field interactions between excitons with the same spins, long-living correlation due to the formation of biexciton state by excitons with opposite spins, and short-memory correlation effects in the continuum of unbound two-exciton states. The DFWM spectra give us the relative contributions of these effects and the upper limit for the time of the exciton-exciton correlation in the unbound two-exciton continuum. The obtained results reveal the basis of the cavity polariton scattering model for the DFWM processes in high-Q GaAs microcavity.
Importance of the effects associated with the exciton-exciton interaction in the nonlinear optical response at the fundamental band edge has been revealed in a number of experiments in $`\chi ^{(3)}`$-regime. These effects, which originate from the electromagnetic Coulomb interaction between photo-generated carriers and Pauli exclusion principle, are referred to as four particle correlation effects . Extensive effort has been devoted to investigate such correlations in semiconductors by four-wave mixing spectroscopy, which gives us a unique opportunity to evaluate limits of applicability of the fermionic mean field theory . However, under actual experimental conditions, the nonlinear optical measurements are affected by a number of intrinsic and extrinsic effects, which are caused by non-electronic degrees of freedom. These effects can not be discriminated in a nonlinear optical experiment in wide spectral region, preventing the direct comparison between theory and experiment.
However, the situation becomes more transparent when we restrict ourselves to the spectral window close to the exciton resonance, where the effects of four particle correlation can be classified in terms of spin dependent interaction between excitons . In particular, a simple model , referred to as the weakly interacting boson (WIB) model, in which the interaction between $`1s`$ excitons is described by using two parameters to account for the interaction between excitons with the same and opposite spins, has allowed us to reproduce the overall features of the polarization-sensitive DFWM spectra.
The remarkable efficiency of the WIB model in the describing the polarization-sensitive DFWM measurements in time and frequency domain has made it a promising and handy scheme to study the correlation effects by methods of nonlinear spectroscopy. However, the underlying physics and, especially, the role of second,- and higher-order in exciton-exciton interactions effects , still remain unclear because the model does not account for memory effects. These effects are responsible for the temporal evolution of the $`2\omega `$-coherence within the continuum of the unbound two-exciton states and the long-living correlation due to the existence of biexciton, which has also been found to be important in FWM at semiconductor band edge . Note, that the evolution equation for the excitonic polarization with account for the biexciton state has been obtained in as a natural extension of the WIB model. Similar semi-phenomenological treatment, which ignores the memory effects due to the unbound two-exciton continuum, can be developed by starting from the fermionic description of the excitonic nonlinearity . The study of the effects of the bound and unbound two-exciton states in the third-order optical response is especially interesting in GaAs system, where - in strong contrast to the wide gap semiconductors such as CuCl or ZnSe \- the biexciton binding energy is the order of the exciton linewidth .
In order to elucidate the role of the memory effects at the fundamental band edge, a special attention should be paid to the choice of the material for the nonlinear-optical measurements. Specifically, the experiment should be performed in a system where effects associated with the finite exciton population and inhomogeneity are minimized. In this condition the effects of the exciton population can be minimized by performing pump-probe experiment at large detuning from the excitonic resonance. In this condition the polarization-sensitive shift of the excitonic resonance, which is referred to as the optical Stark effect, is the signature of the four-particle correlation . At resonance condition, the exciton-cavity coupled system is a well suited candidate to observe the signature of the four-particle correlation. This is because strong coherent coupling between exciton and photon in the high-Q microcavity, which leads to the formation of cavity polaritons and suppresses incoherent effects in the DFWM signal . In this paper, we show that the study of polarization-sensitive DFWM spectra in GaAs/AlGaAs QW embedded in the high-Q microcavity allows us to discriminate the memory effects. By comparing the results of the experiment and theory, we show that the correlation memory time of excitons in GaAs is very short justifying the description of the DFWM process, which is based on cavity polariton scattering. We also obtain the relative contributions from the bound and unbound two-exciton states to the nonlinear optical response.
The semiconductor microcavity investigated in this work has a single 12-nm-thick GaAs quantum well at the antinode of a $`\lambda /2`$-planar microcavity, consisting of 22 and 14.5 pairs of distributed Bragg reflectors for the bottom and topside respectively. All the measurements were performed at 13 K. The linear reflection spectrum shows that the normal mode splitting at zero detuning (4.3 meV) is larger than the linewidth of either exciton or cavity mode (1.5 meV). The DFWM signal was measured in (xxx)-, (xyy)-, (+++)-, (x++)- and (x+-)-configurations, which are abbreviated by the polarizations of the pump, test and signal beams, respectively, in the self-pumped phase conjugation geometry . We used the tunable picosecond pulses (pulse width of 1.9 psec and spectral width of 0.7 meV) from a Kerr-lens mode-locked Ti:Sapphire laser with a 76 MHz repetition rate. In order to ensure that the measurements were performed in the $`\chi ^{(3)}`$-regime, we examined excitation-power dependence of the DFWM signal and found that it is proportional to $`I_{pump}^2I_{test}`$. The detailed description of the experiment can be find in .
The DFWM spectra at zero exciton-cavity detuning are presented in Fig.1a for different polarization configurations . In the (x++)\- and (+++)-configurations, the spectra experiment show nearly the same intensities for upper and lower polaritons. In the (x+-)\- configuration, where the DFWM signal is dominated by the cavity polariton scattering is due to two-exciton states with zero angular momentum, the signal at lower mode is found to be 2.5 times of that at upper mode. The DFWM spectra in (xxx)\- and (xyy)-configurations show a switching between upper and lower modes.
In order to explain these results, we need to examine how the memory effects in the exciton interaction manifest themselves in different polarization configuration. Following we consider the resonant excitation only and start from the equation of motion for the normalized complex amplitudes of the right- and left-circular components of the excitonic polarization at the frequency $`\omega `$, $`p_\pm =<b_\pm >/(Vv_e)^{1/2}`$, where $`b_\pm `$ is the exciton annihilation operator, $`V`$ and $`v_e`$ are the crystal and exciton volume respectively, subscripts $`\mathrm{"}\pm \mathrm{"}`$ label right- and left-circular components. The exciton volume is defined from the conventional relationship between the exciton and interband dipole moments: $`\mu _{ex}/\mu =(V/v_e)^{1/2}`$. The evolution equation for $`p_\pm `$ can be presented in the following form:
$`i{\displaystyle \frac{p_\pm }{t}}+\mathrm{\Delta }p_\pm =`$ $`(1C|p_\pm |^2)\mathrm{\Omega }_\pm +`$ (2)
$`p_\pm ^{}{\displaystyle _0^{\mathrm{}}}F(\tau )p_\pm ^2(t\tau )e^{2i\mathrm{\Delta }\tau }๐\tau p_{}^{}{\displaystyle _0^{\mathrm{}}}G(\tau )p_+(t\tau )p_{}(t\tau )e^{2i\mathrm{\Delta }\tau }๐\tau `$
Here $`\mathrm{\Omega }_\pm =\mu E_\pm /\mathrm{}`$ are the Rabi frequencies, which correspond to the right- and left-circular components of the electric field $`E_\pm `$ of the light wave at the QW, $`\mathrm{\Delta }=\omega _e\omega i\gamma `$, $`\omega _e`$ and $`\gamma `$ are the exciton frequency and dephasing rate, respectively; $`C>0`$ is the phase space filling (PSF) constant ; $`F(\tau )`$ and $`G(\tau )`$ are memory functions, which account for both instantaneous and retarded parts of the exciton-exciton interaction.
Since in our experiment, the pulse is longer than the exciton dephasing time, we can consider the steady state approximation ignoring the time dependence of slowly varying amplitudes $`p_\pm `$ and $`\mathrm{\Omega }_\pm `$ in Eq. (2). By using Eq. (2) and the evolution equation for the electric field at the QW , the steady-state amplitudes of the third-order polarization at the frequency $`\omega `$, which is responsible for the DFWM signal in the phase conjugated geometry, can be presented in the following form:
$$p_\pm ^{(3)}=A\{E_{\pm ,pump}^2[C\mathrm{\Delta }+_0^{\mathrm{}}F(\tau )e^{2i\mathrm{\Delta }\tau }๐\tau ]+E_{,pump}^2_0^{\mathrm{}}G(\tau )e^{2i\mathrm{\Delta }\tau }๐\tau \}E_{\pm ,test}^{}$$
(3)
where $`A`$ accounts for resonance enhancement of the electric field in the microcavity, $`E_{\pm ,pump}`$ and $`E_{\pm ,test}`$ are amplitudes of the electric field associated with the pump and test beams, respectively, at the QW.
In order to show how the memory effects manifest themselves in the nonlinear response, we separate the memory function $`F(\tau )`$ in terms of instantaneous (mean field) part given by $`\varphi >0`$, and retarded (correlation) part , which is given by $`\mathrm{\Phi }(\tau )`$: $`F(\tau )=\varphi \delta (\tau )\mathrm{\Phi }(\tau )`$. The mean field parameter $`\varphi `$ is of the first order in exciton-exciton interaction and describes the interaction between excitons with same spins and zero center-of-mass momentum , while $`\mathrm{\Phi }(\tau )`$ accounts for correlation effects of the second- and higher-order of the interaction between two excitons. The memory function $`G(\tau )`$ accounts for the effects arising from the interaction between excitons with opposite spins. In this case, the first-order in exciton-exciton interaction term vanishes and this memory function contains correlation effects of the second- and higher order in interaction between two excitons. In this configuration, there exists a bound state of two excitons. Correspondingly, we separate $`G(\tau )`$ in terms of contribution from the unbound two-exciton part given by $`\mathrm{\Psi }(\tau )`$ and biexciton part given by $`\psi e^{i\omega _B\tau }`$, where $`\omega _B`$ is the biexciton binding energy: $`G(\tau )=\mathrm{\Psi }(\tau )i\psi e^{i\omega _B\tau }`$ .
In the high-Q microcavity, the strong coupling between excitons and photons produces polariton modes, which dominate both linear and DFWM spectra. These modes are referred to as lower and upper cavity polaritons. At zero exciton-cavity detuning their frequencies are $`\omega _{\alpha ,\beta }=\omega _eg`$, respectively, where $`g=(2\pi \omega \mu ^2/\mathrm{}nv_e)^{1/2}`$ is the energy of the dipole coupling between exciton and photon and $`n`$ is the refractive index. The intensities of the polarization-sensitive DFWM signal at frequencies of the lower and upper polaritons for different polarization configurations can be obtained from (3) as follows: $`I_{\alpha ,\beta }^{xxx}|R+W\pm (Cg+\delta R+\delta W)|^2`$, $`I_{\alpha ,\beta }^{xyy}|RW\pm (Cg+\delta R\delta W)|^2`$, $`I_{\alpha ,\beta }^{+++}|R\pm (Cg+\delta R)|^2`$ and $`I_{\alpha ,\beta }^{x+}|W\pm \delta W|^2`$, where
$`R=\varphi {\displaystyle _0^{\mathrm{}}}\mathrm{\Phi }(\tau )e^{2\gamma \tau }\mathrm{cos}2g\tau d\tau `$ (4)
$`\delta R=i{\displaystyle _0^{\mathrm{}}}\mathrm{\Phi }(\tau )e^{2\gamma \tau }\mathrm{sin}2g\tau d\tau `$ (5)
$`W={\displaystyle _0^{\mathrm{}}}\mathrm{\Psi }(\tau )e^{2\gamma \tau }\mathrm{cos}2g\tau d\tau +{\displaystyle \frac{(2ig+\omega _B)\psi }{(2i\gamma +\omega _B)^24g^2}}`$ (6)
$`\delta W=i{\displaystyle _0^{\mathrm{}}}\mathrm{\Psi }(\tau )e^{2\gamma \tau }\mathrm{sin}2g\tau d\tau +{\displaystyle \frac{2g\psi }{(2i\gamma +\omega _B)^24g^2}}`$ (7)
Normal mode splitting $`2g`$ is the major spectral characteristic of the strongly coupled exciton-cavity system and, correspondingly, the role of the memory effects in the excitonic nonlinear response is determined by its ratio to the spectrum width of the memory functions. In order to clarify the role of the memory effects in the DFWM spectra, we first examine the long-time memory limit case, i.e. $`2g>>\tau _c^1`$, where $`\tau _c`$ is the correlation time of the memory functions $`\mathrm{\Phi }(\tau )`$ and $`\mathrm{\Psi }(\tau )`$. By simplifying Eq. (7) with account for $`2g\tau _c>>1`$ and $`g>>\gamma ,\omega _B`$ one can arrive at the following equations for the polariton intensities: $`I_{\alpha ,\beta }^{+++}|\varphi \pm Cg|^2`$, $`I_{\alpha ,\beta }^{xxx}|\varphi \pm (Cg\psi /2g)|^2`$ and $`I_{\alpha ,\beta }^{xyy}|\varphi \pm (Cg+\psi /2g)|^2`$. Since $`\varphi ,\psi >0`$ and $`\gamma ,\omega _B<2g`$ one may see that $`I_\alpha ^{xyy}>I_\alpha ^{+++}`$ (intensity of the lower polariton in (xyy)-configuration is higher than that in ($`+++`$)-configuration) and $`I_\beta ^{xxx}>I_\beta ^{+++}`$ (intensity of the upper polariton in (xxx)-configuration is higher than that in ($`+++`$)-configuration). However, it can be clearly observed from the spectra 1n Fig. 1a, that such a conclusion contradicts to the experimental results making assumption $`2g\tau _c>>1`$ is invalid.
Therefore, our experimental findings invoke the condition $`2g<\tau _c^1`$. In such a case, we can substitute $`e^{2\gamma \tau }cos2g\tau 12\gamma \tau 2g^2\tau ^2+\mathrm{}`$ and $`e^{2\gamma \tau }sin2g\tau 2g\tau +\mathrm{}`$, and neglect $`\delta R`$ in comparison with $`Cg`$. Similarly, at $`\omega _B<2g`$ one can estimate $`W_0^{\mathrm{}}\mathrm{\Psi }(\tau )๐\tau `$ and $`\delta W\psi /2g`$. These gives $`I_{\alpha ,\beta }^{xxx}|R+W\pm (Cg+\delta W)|^2`$, $`I_{\alpha ,\beta }^{xyy}|RW\pm (Cg\delta W)|^2`$, $`I_{\alpha ,\beta }^{+++}|R\pm Cg|^2`$ and $`I_{\alpha ,\beta }^{x+}|W\pm \delta W|^2`$ ensuring $`I_\alpha ^{+++}>I_\alpha ^{xyy}>I_\alpha ^{xxx}`$ and $`I_\beta ^{+++}>I_\beta ^{xxx}>I_\beta ^{xyy}`$. Note that since we observe $`I_\alpha ^{x+}>I_\beta ^{x+}`$, the following relationship holds: $`W<\delta W<0`$. The observed difference in the intensities of the upper and lower polaritons in the (x+-)-configuration originates from the bound two-exciton state. The ratio $`I_\alpha ^{x+}/I_\beta ^{x+}2.5`$ obtained in the experiment is consistent with the followed from the sum rule theoretical estimation $`I_\alpha ^{x+}/I_\beta ^{x+}1+2\omega _B/g`$, for typical GaAs biexciton binding energy . The calculated DFWM spectra with account for the bound and unbound two-exciton states are presented in Fig. 1b for $`(W):\delta W:R:Cg=0.7:0.23:0.12:1`$. We would like to note here, that our estimation $`\tau _c<(2g)^1`$ is consistent with the results of the calculation of the memory functions within the 1D-Hubbard model framework .
With account for $`\delta R<<Cg`$ and $`|\delta W|<|W|`$, the obtained result returns the prediction of the polariton scattering model , which has allowed us obtain Eq. (3) with the frequency independent $`W`$ and $`R`$. In this model, these parameters account for the attraction and repulsion between excitons with opposite and same spins, respectively, in the phenomenological WIB Hamiltonian . The experimental spectra at both zero (see Fig. 1a) and arbitrary detuning for the above mentioned polarization configurations have been explained by the following relationship between the parameters of the polariton scattering model: $`(W):R:Cg=0.75:0.1:1`$ . This has allowed us to conclude that the attractive interaction between excitons and the PSF effect dominate in the DFWM process in the high-Q microcavity. Apparently the polariton scattering model failed to explain the difference in $`I_\alpha ^{x+}`$ and $`I_\beta ^{x+}`$, which is due to the biexciton effect. Nevertheless, the parameters, which has been obtained in , coincide with our present estimations, because the biexcitonic effects do not affect significantly the spectra in (xxx)\- and (xyy)-configurations. The relatively small value of the parameter $`R`$ obtained in the experiment is due to nearly cancellation of the first- and higher-order in the exciton-exciton interaction contributions to the resonant third-order susceptibility. This cancellation has also been discussed in in terms of constraints, which are imposed on the spectral density of the memory functions by the sum rules.
In conclusion, we formulate the resonant DFWM results in terms of exciton memory functions and show that the polarization-sensitive DFWM spectra give us an important information on the memory effects in exciton-exciton interaction. The intensity of the DFWM signal in the strongly coupled exciton-cavity system is determined by both short-memory correlation in the unbound two-exciton continuum and long-memory correlation associated with the biexciton state. Both these effects give the second- and higher-order in exciton-exciton interaction contributions to the resonance optical nonlinearity, while the mean field, instantaneous contribution, which is of the first-order in the exciton-exciton interaction, is nearly canceled. By comparing the results of the experiment and theory in various polarization configurations, we estimate the upper limit of the correlation time of the memory functions, $`\tau _c<<900`$ fs, which describe the $`2\omega `$-coherence due to the continuum of the unbound two-exciton states. This also allows us to show that the relative contribution to the resonant third-order susceptibility from the biexciton in GaAs is about 30 percent. We show that the short memory time of the exciton-exciton interaction permits to describe the coherent optical response of the excitons in the high-Q semiconductor microcavity in terms of the cavity polariton scattering model, which should be extended to account for the biexciton state.
Discussions with Professor L. J. Sham, Professor A. Shimizu and Dr J. Inoue are duly acknowledged. The microcavity has been provided by Professor H. Sakaki and Dr T. Someya. We are also grateful to Dr C. Ramkumar, R. Shimano and T. Aoki for helpful discussions. M.K.-G. is partially supported by a grant-in-aid for COE Research from the Ministry of Education, Science, Sports, and Culture of Japan. |
warning/0003/astro-ph0003301.html | ar5iv | text | # Physical conditions in broad and associated narrow absorption-line systems toward APM 08279+5255 Based on observations collected at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation.
## 1 Introduction
The gravitationally lensed high-redshift Broad Absorption Line (BAL) QSO APM 08279+5255 has been given tremendous interest since its discovery by Irwin et al. (1998) as it is one of the most luminous objects in the universe even after correction for gravitational amplification. Based on the position of the emission lines, Irwin et al. (1998) derived a redshift $`z_{\mathrm{em}}`$ = 3.87. A probably better estimate of the systematic redshift comes from the detection of CO(4โ3) emission at $`z_{\mathrm{em}}`$ = 3.9114$`\pm `$0.0003 by Downes et al. (1999).
Imaging of the field reveals two main components (Irwin et al. 1998, Ledoux et al. 1999) separated by 0.378$`\pm `$0.001 arcsec as measured on HST/NICMOS data (Ibata et al. 1999) and of relative brightness $`f_\mathrm{B}/f_\mathrm{A}`$ = 0.773$`\pm `$0.007. The HST images reveal the presence of a third object C with $`f_\mathrm{C}/f_\mathrm{A}`$ = 0.175$`\pm `$0.008, located in between A and B and almost aligned with them. The PSF fits on the three objects are consistent with the three components being point-sources and the colors are similar within the uncertainties suggesting that C is a third image of the quasar (Ibata et al. 1999, Egami et al. 1999). A high-resolution high signal-to-noise ratio spectrum of APM 08279+5255, covering the wavelength range 4400โ10000ร
was obtained using the Keck telescope and made available to the Astronomy community for analysis (Ellison et al. 1999a,b). This spectrum, though complicated by the combination of light traveling along three different sight lines, is a unique laboratory for studying the intervening and associated absorption systems.
It is well known that the origin of associated systems (systems with $`z_{\mathrm{abs}}`$ $``$ $`z_{\mathrm{em}}`$) cannot be inferred directly from their position in the spectrum. Indeed, absorption can arise from (i) gas ejected by the central engine at velocity as high as 60000 km s<sup>-1</sup> and nonetheless physically located very close to the source of ionizing photons (e.g. the $`z_{\mathrm{abs}}`$= 2.24 โmini-BALโ towards Q 2343+125 at $`z_{\mathrm{em}}`$ = 2.515; Hamann, Barlow & Junkkarinen 1997a) or from (ii) gas associated with the host-galaxy or with members of a galaxy cluster surrounding the quasar. The distinction can be made in terms of physical properties. The systems belonging to the first class are characterized by high metal enrichment, high-ionization parameters, broader line profiles, partial coverage and time variability (Barlow et al. 1992, Petitjean et al. 1994, Savaglio et al. 1994, Hamann 1997, Hamann et al. 1997b, Barlow & Sargent 1997, Ganguly et al. 1999, Papovich et al. 1999). The second class of absorber is characterized by classical properties of intervening systems such as low metallicities (typically 0.01 to 0.1 of solar) and undisturbed kinematics.
In this study we investigate the nature and physical properties of $`z_{\mathrm{abs}}`$$``$ $`z_{\mathrm{em}}`$ systems toward APM 08279+5255. In Section 2, we describe the data and the grids of photoionization models we use to infer ionizing conditions in the absorbing gas. We analyse a probably intervening metal line system very close to the emission redshift in Section 3. The nature of narrow-line systems with low-ionization conditions are investigated in Section 4. Section 5 describes the high-ionization narrow-line systems. In Section 6, we analyse the nature of the BAL outflow and in Section 7 we suggest the presence of โline-lockingโ among narrow components in the BAL flow. A summary is given in Section 8.
## 2 Data and photoionization models
A spectrum of the $`z_{\mathrm{em}}`$ = 3.91 quasar APM 08279+5255 was obtained with the HIRES echelle spectrograph at the 10m Keck-I telescope (Ellison et al. 1999a,b). This data was made public together with a low-resolution spectrum of the quasar and a high-resolution spectrum of a standard star. We have corrected the high-resolution spectrum of APM 08279+5255 for small discontinuities in the continuum, which are probably due to the inappropriate merging of different orders. These discontinuities have been recognized by comparing the high and low-resolution spectra. The latter has also been used for normalization of the high-resolution data. Atmospheric absorption features were identified from the standard star spectrum. We have measured the final spectral resolution by fitting the narrow atmospheric absorption lines which are free of blending. We find $`FWHM`$ $``$ 8 km s<sup>-1</sup> ($`b`$ $``$ 4.8 km s<sup>-1</sup>) at 6900 ร
, $`R`$ = 37500, and use this value throughout the paper.
Grids of photoionization models using the code Cloudy (Ferland 1996) have been constructed. The cloud is a plane parallel slab of uniform density, solar chemical composition and neutral hydrogen column densities $`N`$(H i) = $`10^{16}`$ cm<sup>-2</sup> (see below) photo-ionized by the QSO radiation. The spectral energy distribution of APM08279+5255 is not known in the UV/X-ray energy range. We use the standard AGN spectrum provided by Mathews & Ferland (1987) and also consider the effect of screening the ionizing spectrum by optically thick clouds to take into account the fact that BALs are X-ray quiet (Green & Mathur 1996). The resulting column densities of various species, for different ionizing spectra, along a line-of-sight perpendicular to the slab are given in Fig. 1.
## 3 The intervening system at $`z_{\mathrm{abs}}`$ = 3.8576
This system is revealed by strong C iv and hydrogen Lyman series lines. It is at higher redshift than the main component of the BAL out-flowing gas. Moderately saturated C iii$`\lambda `$977 and C iv absorptions are seen whereas Si iii$`\lambda `$1206 and Si iv, although present, are weak (see Fig. 2). Both C iv and C iii$`\lambda `$977 profiles suggest the presence of 5 distinct components. Only the highest velocity component shows absorption due to higher hydrogen Lyman series transitions. The expected positions of O vi is in the wavelength range of the very strong O vi BAL.
Column densities estimated from Voigt profile fitting are given in Table LABEL:tab385 for different species. Note that the C iv doublet ratio is consistent with complete coverage of the background source (see below). The neutral hydrogen column density at $`z_{\mathrm{abs}}`$ = 3.8574 is well defined by higher Lyman series absorption lines. If we assume the QSO is the ionizing source we derive log $`U=2.5`$ from the C iii and C iv column densities and the absence of singly ionized species. Models with solar metallicities produce C iii and C iv column densities two orders of magnitude larger than what is observed (see Fig. 1). This clearly suggests that metallicities are of the order of \[C/H\] $``$ 0.01 \[C/H\], similar to what is usually derived for intervening systems. It is likely that this system does not belong to the QSO environment.
## 4 Intrinsic systems with narrow and low-ionization lines
### 4.1 The $`z_{\mathrm{abs}}`$= 3.8931 complex
Absorption profiles of some of the important transitions from this complex are given in Fig. 3. In what follows, we concentrate on the component at $`v`$ $``$ 0 km s<sup>-1</sup>, which shows absorption due to low ionization lines: Al ii, Al iii, Si ii, Si ii\*, C ii and C ii\*. In addition, there are strong absorption lines at the expected positions of N iii$`\lambda `$989 and C iii$`\lambda `$977 with profiles similar to that of C iv and Si iv. It is however very difficult to rule out contamination by intervening Lyman-$`\alpha `$ systems. There is a feature at the expected position of O vi$`\lambda `$1031. However, as O vi$`\lambda `$1037 is heavily blended, it is difficult to establish the presence of O vi in this system. N v absorption is present and weak (see Fig. 3).
#### 4.1.1 Partial covering factor
The flat bottom and the presence of other Lyman series lines show that the Lyman-$`\alpha `$ line is saturated. However, it is apparent on Fig. 3 that there is some residual flux in the core of the Lyman-$`\alpha `$ line. Similar residuals are seen at the bottom of Siiv$`\lambda `$1393 and Civ$`\lambda `$1548. Residual flux is expected in case one of the three lines of sight is not intercepted by the absorbing cloud. However the relative brightnesses of the components are $`f_\mathrm{B}/f_\mathrm{A}`$ $``$ 0.773 and $`f_\mathrm{C}/f_\mathrm{A}`$ $``$ 0.175 (Ibata et al. 1999). Therefore, under the above assumption, at least 10% residual flux is expected, much larger than the 3% observed. This clearly suggests instead that the absorbing gas is located outside the BLR and that it covers only 97% of the background source (BLR+continuum+scattered light). We use the C iv, Si iv and Al iii doublets to estimate the covering factor of these species applying the method described in Srianand & Sankaranarayanan, 1999. The results are shown in Fig. 4. The covering factor of the C iv and Si iv component at $`v`$0 km s<sup>-1</sup> is about 0.95 whereas the corresponding value for Al iii is 0.80.
For simplicity and because the residual is small, we assume complete coverage while estimating the column density of H i. We find log $`N`$(H i) = 15.9, suggesting that the partial Lyman limit detected in the rest frame of the QSO (Irwin et al. 1998) is most certainly not due to this system.
The covering factor of the singly ionized phase can be estimated using the numerous Si ii lines. In particular, the Si ii$`\lambda `$1260 and Si ii$`\lambda `$1304 absorption lines are redshifted beyond the QSO Lyman-$`\alpha `$ emission line. The Si ii$`\lambda `$1260 line is blended with C iv$`\lambda `$1548 at $`z_{\mathrm{abs}}`$= 2.9833 and we use C iv$`\lambda `$1550 to remove the contamination. In the lower panel of Fig. 5, the best Si ii$`\lambda `$1260 fit, after removal of the C iv contribution, is plotted together with the Si ii$`\lambda `$1304 profile appropriately scaled to Si ii$`\lambda `$1260 for a covering factor of $`f_\mathrm{c}=0.8`$. This covering factor is consistent with the value derived from the Al iii doublet. In the top panel of Fig. 5 the Si ii\*$`\lambda `$1309 profile is plotted together with the scaled profile of Si ii\*$`\lambda `$1264 for two values of the covering factor. The best value is $`f0.54`$, however as the Si ii\*$`\lambda 1309`$ is weak, a value of 0.8 is well within the 1$`\sigma `$ error. Thus, the data is consistent with both Si ii and Si ii\* absorptions having similar covering factors $`f_\mathrm{c}0.8`$.
The column densities given in the fourth column of Table LABEL:tab389 are obtained by integrating the optical depth over the velocity range covered by the absorptions after taking into account the covering factors derived above. Note that the C iv column density is unreliable due to blending with other lines as well as contamination of C iv$`\lambda `$1550 by C iv$`\lambda `$1548 at $`z_{\mathrm{abs}}`$= 3.901. Results from single component Voigt-profile fits are given in the second column of the Table LABEL:tab389.
#### 4.1.2 Excitation of the fine-structure levels
The presence of several absorption lines from excited fine-structure levels of Si ii and C ii is a unique opportunity to study the physical properties of the absorbing gas. Two main excitation processes are at play: radiative excitation by an IR radiation field and collisional excitation mainly by electrons.
Let us first consider radiative excitation. In that case,
$$\frac{N(X^{})}{N(X)}=2\left(\frac{\overline{n}_\lambda }{1+\overline{n}_\lambda }\right)$$
(1)
where $`X`$ is either C ii or Si ii and $`\overline{n}_\lambda `$ is given by
$$\overline{n}_\lambda =\frac{I(\nu )\lambda ^3}{8\pi hc}$$
(2)
where $`I(\nu )`$, in $`\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1\mathrm{Hz}^1`$, is the flux of energy density integrated over all directions. From the available IRAS fluxes and the sub-millimeter observations by Lewis et al. (1999), the flux at the excitation energy of Si ii (0.036 eV) in the rest frame of the absorber is 0.1 Jy. Assuming the IR emitting region to be a point, $`z`$ = 3.910, $`H_\mathrm{o}`$ = 75 km s<sup>-1</sup> Mpc<sup>-1</sup> and $`q_\mathrm{o}`$ = 0.5, we obtain,
$$\overline{n}_\lambda =\frac{0.0038}{r(\mathrm{kpc})^2}\left(\frac{1}{k}\right)$$
(3)
where $`r`$ is the distance of the cloud from the ionizing source and $`k`$ is the magnification factor due to gravitational lensing. From Eqs. (1) and (2) we derive,
$$r(\mathrm{kpc})^2=\left[2\frac{N(\mathrm{Si}\mathrm{ii})}{N(\mathrm{Si}\mathrm{ii}^{})}1\right]\times \frac{0.0038}{k}$$
(4)
The IR flux at the excitation energy of C ii (0.008 eV) is 0.016 Jy so that
$$\overline{n}_\lambda (\mathrm{C}\mathrm{ii}^{})=15.18\times \overline{n}_\lambda (\mathrm{Si}\mathrm{ii}^{})$$
(5)
Therefore if radiative excitation is at play, we should observe,
$$\frac{N(\mathrm{C}\mathrm{ii})}{N(\mathrm{C}\mathrm{ii}^{})}=\left[\frac{N(\mathrm{Si}\mathrm{ii})}{N(\mathrm{Si}\mathrm{ii}^{})}+7.09\right]\times 0.063.$$
(6)
As $`N`$(Si ii)/$`N`$(Si ii\*$``$ 3.5, the radiative excitation predicts $`N`$(C ii)/$`N`$(C ii\*$``$ 0.67 when the observed ratio is 0.66. This strongly suggests that indeed excitation by IR radiation is important. For $`k=1`$, we obtain an upper limit on the distance of the cloud from the IR source, $`r`$ $`<`$ 150 pc. If the magnification factor is as high as 90, as suggested by the probable presence of the third image (Ibata et al. 1999, Egami et al. 1999), the distance could be as small as 15 pc.
Let us consider now the case of pure collisional excitation. Note that excitation by hydrogen atoms is unimportant in the case of optically thin clouds. Deexcitation is due to spontaneous emission as well as particle collisions. The necessary relationships to estimate the population ratios as a function of temperature and density are taken from Bahcall & Wolf (1968). For Si ii the collisional excitation rate is, $`3.32\times 10^7T_4^{0.5}exp(413.4/T)`$ cm<sup>3</sup>s<sup>-1</sup>, the collisional deexcitation rate is, $`1.66\times 10^7T_4^{0.5}`$ cm<sup>3</sup>s<sup>-1</sup> and the spontaneous radiative deexcitation rate is $`2.13\times `$ $`10^4`$ s<sup>-1</sup>. For C ii the above rates are respectively, $`6.20\times 10^8T_4^{0.5}`$ $`exp(91.25/T)`$ cm<sup>3</sup>s<sup>-1</sup>, $`3.10\times 10^8T_4^{0.5}`$ cm<sup>3</sup>s<sup>-1</sup> and $`2.36\times 10^6`$ s<sup>-1</sup>. Here, $`T_4`$ is the temperature expressed in units of $`10^4`$ K.
In Fig. 6, are plotted the ratios $`N`$(Si ii\*)/$`N`$(Si ii) and $`N`$(C ii\*)/$`N`$(C ii), of excited to ground state column densities as a function of temperature for three different values of the electronic density. In order to reproduce the observed ratios (indicated by horizontal dashed lines in Fig. 6), the electron density has to be larger than 100 cm<sup>-3</sup> and the temperature greater than 320 K. As can be seen from Fig. 6, there is no stringent constraint on the electron density when the electron temperature of the gas is low. Note that although it would be surprising to find such a small temperature in an optically thin cloud located in the vicinity of the quasar, this is possible if the metallicity of the gas is well above solar (e.g. Petitjean et al. 1994). In case the gas is warm, from the width of the absorption lines, we derive an upper limit on the temperature of 4.6$`\times 10^4`$ K which gives an upper limit on the density of 500 cm<sup>-3</sup>.
Is there any way to choose between the two processes discussed above? The small distance inferred in case of pure radiative excitation together with the low-ionization state of the gas may be at odd with the proximity of a very powerful source of ionizing radiation. From the low spectral resolution spectrum, we estimate the flux at the Lyman limit in the rest frame of the absorber to be 3$`\times 10^{16}`$ erg s<sup>-1</sup>cm<sup>-2</sup>ร
<sup>-1</sup>. This corresponds to a luminosity at the Lyman limit of $`L_\nu =4.28\times 10^{30}`$ erg Hz<sup>-1</sup>s<sup>-1</sup>. Assuming a flat spectrum and integrating $`L(\nu )/h\nu `$ over the energy range 1 to 20 Ryd, we estimate the number of ionizing photons per unit time emitted by the quasar, $`Q`$ = 2$`\times `$10$`{}_{}{}^{57}/k`$, where $`k`$ is the magnification factor. Thus, the distance $`r`$ between the cloud and the quasar is
$$r=\frac{24.3}{\sqrt{(kn_{100}U_2)}}\mathrm{kpc}$$
(7)
where $`n_{100}`$ is the particle density in units of $`10^2`$ cm<sup>-3</sup>, and $`U_2`$ is the dimensionless ionization parameter in units of $`10^2`$. The latter is the ratio of the density of ionizing photons to the hydrogen density. The H i, Si ii and C ii column densities derived in this system are consistent with log $`U`$ $``$2 (see Table LABEL:tab389 and Fig. 1). Even for metallicities larger than 10 $`Z_{}`$, log $`U`$ $`<`$1. As $`n_{100}`$ $`<`$ 5 and $`k`$ $`<`$ 90, we conclude that $`r`$ $`>`$ 350 pc. This lower limit can be reconciled with the distance inferred from the IR excitation if (i) contrary to the UV source, the IR source is not lensed which does not seem to be the case (see Downes et al. 1999) or (ii) the cloud is closer to the IR source than to the UV source. Indeed Downes et al. (1999) have shown that the molecular gas and the dust are located in a nuclear disk of radius 90 to 270 pc. The absorption system could be at 350 pc from the central UV source and still very close to the IR emitting disk.
Let us now assume that the cloud is at a large distance from the quasar and that the excitation is purely collisional. In that case, the electronic density must be larger than 100 cm<sup>-3</sup>. Rapid variations of the density on very small scales could imply variations of the Si ii/Si ii ratio from one line-of-sight to the other. If we apply a typical ionization correction, H i/H<sup>+</sup> = 4$`\times `$10<sup>-3</sup>, to the H i column density derived above (see Table LABEL:tab389), we obtain $`N`$(H) $``$ 2$`\times `$10<sup>18</sup> cm<sup>-2</sup> and assuming $`n`$ $`>`$ 100 cm<sup>-3</sup>, we infer a dimension of the cloud along the line-of-sight of $``$0.006 pc. Reverberation studies of the nearby AGN NGC 5548 suggest that the size of the BLR could be in the range 10 to 20 lt days (e.g. Krolik & Done 1995, Korista et al. 1995). It is also known that the size of the BLR scales with luminosity as, $`L^{0.5}`$. From the UV and optical spectra we estimate that APM 0827+5255 is at least 500/$`k`$ times brighter than NGC 5548. If we apply the luminosity scaling then the radius of the BLR of APM 0827+5255 is of the order of $``$0.3/$`\sqrt{k}`$ pc. Thus in order to cover 97% of the background source, the transverse dimension of the cloud should be $``$50/$`\sqrt{k}`$ times larger than the dimension along the line-of-sight. As $`k`$ could be as large as 100, the two dimensions could be of the same order of magnitude. The cloud is therefore very small.
#### 4.1.3 Photoionization model
In this Section, we discuss the ionization state of the gas in more detail. Fig. 1 shows the results of models with solar metallicity assuming different ionizing spectra, either an unattenuated Mathews & Ferland (1987) spectrum or the same spectrum attenuated by an ionized hydrogen and helium slab with log $`U`$ = โ1 and H i column density of 10<sup>17-18</sup> cm<sup>-2</sup>. The resultant spectra are given in panel (d) of Fig. 1. It is apparent that the absorption is most important at the He ii edge. This mimics the presence of the BAL outflow along the line-of-sight. As an illustration, the number of ionizing photons is reduced by a factor 6 due to the presence of a screen with log $`N`$(H i) = 18.
In models with unattenuated ionizing radiation, the observed $`N`$(Al iii)/$`N`$(Al ii), $`N`$(Si iv)/$`N`$(Si ii) and $`N`$(C iv)/$`N`$(C ii) ratios suggest consistently log $`U=2.0`$, $`2.2`$ and $`2.3`$ and \[Al/H\] = 3.2 \[Al/H\], \[C/H\] = 0.5 \[C/H\] and \[Si/H\] = 1.6 \[Si/H\]. Models with ionizing radiation attenuated by a screen with $`N`$(H i) = $`10^{17}`$ cm<sup>-2</sup> give log $`U=2.2`$, and \[Al/H\] = 1.7 \[Al/H\], \[C/H\] = 0.5 \[C/H\] and \[Si/H\] = 0.9 \[Si/H\].
The most important effect of attenuation is that the N v column density decreases with increasing attenuation due to the paucity of photons of energy larger than the N iv ionization potential. In the case of unattenuated ionizing spectrum and assuming solar abundances, the model predicts log $`N`$(N v)$`>13.2`$ which is at least a factor of 3 larger than the observed value. Attenuation by a slab with neutral column density as large as log $`N`$(H i) = 18 is ruled out, however, as in that case log $`N`$(N v) = 11 instead of 12.7 observed.
In conclusion, if the shape of the ionizing radiation from the QSO is similar to the Mathews & Ferland (1987) spectrum, then the data is consistent with attenuation of the ionizing flux by a screen with log $`N`$(H i$``$ 17. Absolute metallicity is probably very close to solar which is typical of associated absorption systems (e.g. Hamann & Ferland 1993; Petitjean et al. 1994; Petitjean & Srianand 1999, Hamann & Ferland 1999 for a recent review). However there is an indication that Carbon is under-abundant and Aluminum over-abundant compared to Silicon.
### 4.2 The $`z_{\mathrm{abs}}`$= 3.9135 complex
The redshift of this system is very close to the CO emission redshift, $`z_{\mathrm{CO}}`$ = 3.9114 (Downes et al. 1999); the difference is only $``$130 km s<sup>-1</sup> whereas the width of the CO line is 480 km s<sup>-1</sup> and the total width of the saturated part of the associated Lyman-$`\alpha `$ line is 300 km s<sup>-1</sup>. Three distinct sub-systems are seen in Si iv and C iv. The strongest component in this complex has associated absorption from C ii, C iv, Si ii, Si iv, Al ii, Al iii and Lyman-$`\alpha `$ (see Fig. 7). Unlike what is observed at $`z_{\mathrm{abs}}`$= 3.8931, the Lyman-$`\alpha `$ line goes to zero indicating complete coverage. This is confirmed by the Si iv and Al iii doublets. Using the Lyman series, we estimate the neutral hydrogen column density, log $`N`$(H i) = 17.0$`\pm `$0.14 (see Table LABEL:tab391). This is consistent with this system being responsible for the partial Lyman limit absorption seen in the low-resolution spectrum of APM 08279+5255. The Si iv and C iv doublets are well fitted by a model with three components; all the absorption lines due to singly ionized species are fitted with a single component. N v and O vi are clearly absent. There is no absorption from Si ii\* down to a 2$`\sigma `$ upper limit of log $`N`$(Si ii\*$`<`$ 11.9. The C ii\* line is redshifted at the same position as C iv$`\lambda `$1548 at $`z_{\mathrm{abs}}`$= 3.2388. We estimate a 2$`\sigma `$ upper limit on $`N`$(C ii\*) after subtraction of the C iv profile, log $`N`$(C ii\*$`<`$ 12.7. The ratio of C ii\* to C ii in this system is at least a factor 4 smaller than in the $`z_{\mathrm{abs}}`$= 3.8931 system discussed previously. This suggests that this cloud has to be farther away from the IR source than the $`z_{\mathrm{abs}}`$ = 3.8931 cloud. For electron temperatures less than 3$`\times 10^4`$ K the observed limit also suggests that the electronic density is less than 20 cm<sup>-3</sup>.
The $`N`$(Al ii)/$`N`$(Al iii) column density ratio in the two systems at $`z_{\mathrm{abs}}`$= 3.8931 and 3.9135 can be used to compare their ionization states. $`N`$(Al ii)/$`N`$(Al iii) = 0.40 at $`z_{\mathrm{abs}}`$= 3.9135 while the allowed range at $`z_{\mathrm{abs}}`$= 3.8931 is 0.06โ0.10 (after taking into account covering factor effects). Note that the $`N`$(C ii)/$`N`$(C iv) ratio is also larger at $`z_{\mathrm{abs}}`$= 3.9135. From Fig. 7 of Petitjean et al. (1994) it is clear that when $`N`$(H i$`<`$ 10<sup>17</sup> cm<sup>-2</sup>, $`N`$(Al ii)/$`N`$(Al iii) is approximately independent of the neutral hydrogen column density (for a given ionizing spectrum) and depends only on the ionization parameter. Therefore, the ionization parameter of the $`z_{\mathrm{abs}}`$= 3.9135 system is slightly smaller than that of the $`z_{\mathrm{abs}}`$= 3.8931 system.
To discuss this system in more detail, we can use the models of Fig. 1b, scaling the column densities with abundances. The $`N`$(Al iii)/$`N`$(Al ii), $`N`$(C ii)/$`N`$(C iv) and $`N`$(Si iii)/$`N`$(Si iv) ratios suggest, log $`U=2.83,2.6`$ and $`2.3`$ respectively. The range of ionization parameters reflects the uncertainties. If we use the mean ionization parameter, log $`U=2.5`$, the model suggests \[C/H\] = 0.02 \[C/H\], \[Si/H\] = 0.01 \[Si/H\] and \[Al/H\] = 0.23 \[Al/H\]. Though the average abundance in this system is similar to intervening systems at large redshift, it seems that Aluminum is largely enhanced compared to Carbon and Silicon. Such peculiar abundance pattern has already been noted by Ganguly et al. (1999).
All this suggests that whereas the $`z_{\mathrm{abs}}`$= 3.8931 system must be located within 200 pc from the QSO and ejected at a velocity larger than 1000 km s<sup>-1</sup>, the $`z_{\mathrm{abs}}`$= 3.9135 system must be farther away and part of the host-galaxy.
## 5 High-ionization narrow-line systems
In this section we study the high-ionization narrow-line associated systems, defined as well detached systems with Lyman-$`\alpha `$ apparent optical depths much smaller than that of O vi and N v (see Fig. 8). There are two main complexes at $`z_{\mathrm{abs}}`$$``$3.90 and $`z_{\mathrm{abs}}`$$``$3.917. We concentrate on well defined strong components.
### 5.1 $`z_{\mathrm{abs}}`$= 3.8997
Column densities are estimated using the apparent optical depth method (Savage & Sembach 1991). The results are given in Table LABEL:tabhigh together with the velocity intervals over which the column density is estimated. The residual intensities in both components of the N v doublet and the O vi$`\lambda `$1031 line are consistent with complete coverage. Unlike O vi, N v lines are not heavily saturated. As can be seen from Fig. 8, the C iv$`\lambda `$1550 component is heavily affected by atmospheric absorption. We estimate a lower limit on the C iv column density using the C iv$`\lambda `$1548 profile and assuming complete coverage. There is a strong absorption feature at the expected position of C iii$`\lambda `$977 , the profile of which is very similar to that of C iv. As we cannot rule out contamination by intervening Lyman-$`\alpha `$ absorption, the C iii column density should be considered as an upper limit. As can be seen from Fig. 8, the associated Lyman-$`\alpha `$ absorption is very weak in this component. In addition, this line falls in the wavelength range over which N v BAL absorption is present and coincides with the expected position of the N v$`\lambda `$1242 absorption from a complex at $`z_{\mathrm{abs}}`$= 3.795, which has ubiquitous absorption due to C iv and Si iv (see also next Section). Therefore the H i column density given in Table LABEL:tabhigh is most certainly an upper limit.
### 5.2 $`z_{\mathrm{abs}}`$= 3.9010
This component shows strong O vi absorption with profiles consistent with complete coverage. The Lyman-$`\alpha `$ line coincides with the expected position of the N v$`\lambda `$1242 absorption from the complex at $`z_{\mathrm{abs}}`$= 3.795. Thus, the H i column density given in Table LABEL:tabhigh is an upper limit. From Fig. 8, it can be seen that N v$`\lambda `$1242 from this system coincides with N v$`\lambda `$1238 at $`z_{\mathrm{abs}}`$= 3.9170. Thus we estimate the N v column density using the N v$`\lambda `$1238 line only. As noted before, C iv$`\lambda `$1550 is heavily affected by atmospheric absorption; C iv$`\lambda `$1548 coincides with C iv$`\lambda `$1550 at $`z_{\mathrm{abs}}`$= 3.9135 (see Figs. 7,8) preventing us to estimate the C iv column density in this system. There is a strong line at the expected position of C iii$`\lambda `$977. We estimate an upper limit of the C iii column density assuming complete coverage.
### 5.3 $`z_{\mathrm{abs}}`$= 3.9170
The saturated O vi lines have no residual flux so the absorbing cloud completely covers the background source. The O vi$`\lambda `$1037 line is blended with a strong line most probably due to an intervening cloud. The lower limit on the O vi column density is estimated using the O vi$`\lambda `$1031 line only. As discussed before the N v$`\lambda `$1238 line is blended with the N v$`\lambda `$1242 line at $`z_{\mathrm{abs}}`$= 3.9010 and the N v$`\lambda `$1242 line is blended with Mg ii$`\lambda `$2796 at $`z_{\mathrm{abs}}`$= 1.18. It is therefore difficult to estimate the N v column density accurately. We obtain an upper limit using the N v$`\lambda `$1238 profile. Though the C iv absorption is quite strong, both lines of the doublet are badly blended with atmospheric features.
### 5.4 Modeling
The column densities derived in the three systems are given in Table LABEL:tabhigh. The observed column densities of H i, C iv, N v and O vi are typical of associated absorption systems (Petitjean et al. 1994, Hamann 1997, Petitjean & Srianand 1999).
We run photoionization models using the โOptimizeโ command available in โCloudyโ to derive the best values for the ionization parameter and heavy element abundances. For the Mathews & Ferland (1987) ionizing spectrum we obtain log $`U=0.87`$ and $`Z=180Z_{}`$ for the $`z_{\mathrm{abs}}`$= 3.8997 system and log $`U=0.74`$ and $`Z`$(O,N) = 14$`Z_{}`$ for the $`z_{\mathrm{abs}}`$= 3.9010 system.
However, it is known that BALQSOs are X-ray quiet with optical to X-ray spectral index $`\alpha _{\mathrm{OX}}1.9`$ (Green & Mathur, 1996) while the Mathews & Ferland spectrum has $`\alpha _{\mathrm{OX}}=1.40`$. If we assume the spectra seen by the absorbing clouds to be steeper than the Mathews & Ferland spectrum then we obtain $`Z`$ = 5, 1.4, 0.5, 0.4 $`Z_{}`$ and $`U`$ = 0.3, 0.7, 1.2, 1.3 for $`\alpha _{\mathrm{OX}}`$ = 1.5, 1.7, 1.9 and 2.1 respectively for the $`z_{\mathrm{abs}}`$= 3.8997 system. The reason for this is that the number of photons available for ionization of C iv, N v and O vi decreases when $`\alpha _{\mathrm{OX}}`$ increases. Therefore, it would be of first interest to measure $`\alpha _{\mathrm{OX}}`$ for this quasar.
## 6 The BAL flow
The maximum observed ejection velocity of the C iv absorption is $`12000`$ km s<sup>-1</sup> relative to the systemic redshift $`z_{\mathrm{em}}`$ = 3.911. The BAL is of high-ionization: the O vi absorption is much stronger than the C iv one; N v is also conspicuous; weak Si iv absorption is seen in only a few components (especially those responsible for the strong C iv troughs at $`9670`$ km s<sup>-1</sup> and $``$4670 km s<sup>-1</sup>). The structure of the absorption profiles due to the out-flowing gas is complex as it breaks into components of different widths, depths and covering factors (see Fig. 9).
Weymann et al. (1991) have shown that the mean spectrum of BALQSOs shows two distinct and localized minima separated by about the velocity splitting of N v and Lyman-$`\alpha `$ (i.e.$`5900`$ km s<sup>-1</sup>) located near velocities $``$4700 and $``$10500 km s<sup>-1</sup> relative to the emission redshift. Korista et al. (1993) have found that 22% of BALQSOs do show double troughs. These features suggest that radiative acceleration is important to generate the flows and the observed profiles are due to line-locking (Arav 1996 and references therein).
In APM 08279+5255, the two strongest troughs in the C iv profile are separated by $`4900`$ km s<sup>-1</sup>(see Figs. 9 and 9), smaller than the Lyman-$`\alpha `$โN v velocity splitting. In between these two absorption troughs there are weak C iv absorption lines (resolved even in the low-resolution spectrum). From the shape of the red wing of the C iv emission line it is also apparent that the flux in the blue wing of the C iv emission line at $`7461`$ร
is very close to the expected unabsorbed value (see Fig. 10). This suggests that the two strong troughs are well detached. It is very difficult to produce such a profile in models of out-flowing gas where the double structures are created by radiative processes which produce clustering in the velocity space keeping all absorbing atoms at the same physical location (see the fits to double troughs in Fig. 11 of Arav 1996). We thus believe that the two strong troughs are due to physically distinct absorbing components which are similar to mini-BAL systems seen in a few QSOs (e.g. Petitjean & Srianand 1999).
In what follows we consider smaller velocity intervals and investigate the profiles of different absorptions in more detail. The velocity are taken with respect to $`z_{\mathrm{em}}`$ = 3.911.
### 6.1 The broad C iv absorption at $`v12400`$ km s<sup>-1</sup>
The two broad features at 7295 ร
and 7305 ร
correspond mostly to the two lines of a C iv doublet at $`z_{\mathrm{abs}}`$ $``$ 3.712 ($`v12400`$ km s<sup>-1</sup> in Fig. 9). The profiles do not match exactly however. It is clear from Fig. 11 that there is an extra absorption in the C iv$`\lambda `$1550 profile at $`160`$ km s<sup>-1</sup> from the center of the line. This absorption is tentatively identified as C iv$`\lambda `$1548 at $`z_{\mathrm{abs}}`$$``$ 3.717, the corresponding broad, albeit weak, C iv$`\lambda `$1550 line is blended with the narrower Fe ii$`\lambda `$2600 complex from a Mg ii system at $`z_{\mathrm{abs}}`$ = 1.81. Most of the narrow absorption features seen in the C iv$`\lambda `$1548 and C iv$`\lambda `$1550 profiles are due to atmospheric absorptions.
We believe, this absorption structure is real because of the good match between the profiles of the C iv lines in the red wings (see Fig. 11), and because the lines are located very near the centre of an echelle order and thus cannot be an artifact of imperfect order merging. This system is well detached from the rest of the BAL and therefore the continuum can be fitted accurately. The C iv absorption is most probably saturated as the residual intensities in both lines are similar. However they are equal to 0.87 which clearly indicates that the covering factor is surprisingly small, $`0.13`$. The C iv profiles, in the velocity range not affected by the narrow atmospheric absorption lines between $``$100 km s<sup>-1</sup> and 100 km s<sup>-1</sup>, can be fitted with a covering factor of 0.12 (see top panel in Fig. 11). The typical error in the covering factor is 0.02 (as the rms in the normalised continuum over this wavelength range is 0.02). From Fig. 10 it is apparent that this absorption, does not occur on top the C iv emission line. Thus the low coverage is most likely due to the fact that the absorbing gas does not cover one of the images of the very small continuum source. If the absorbing gas covers only image C, then we expect a covering factor of 0.09. The consistency with the observed value is an additional argument in favor of object C being the third image of the gravitationally lensed quasar. If true, this implies that a strong C iv absorption, with very little residual intensity, is expected in the spectrum of image C whereas very weak or no absorptions are expected in the spectra of A and B. This can be probed with HST/STIS observations.
The width of the absorption lines is $`200`$ km s<sup>-1</sup>. This system must be considered as a โmini-BALโ as the C iv profile is smooth and broader than a typical intervening system but narrower than a typical BAL outflow (Barlow et al. 1997). Si iv absorption is not detected and O vi and N v are redshifted at the same position as other strong lines.
### 6.2 Absorptions in the range $`11000<v<9000`$ km s<sup>-1</sup>
In addition to the broad absorption seen in this wavelength range, there are a few narrow absorption lines identified as C iv doublets.
#### 6.2.1 Narrow lines at $`v10030`$ km s<sup>-1</sup>
The system at $`v10030`$ km s<sup>-1</sup> is defined by C iv, O vi, N v and Si iv narrow absorption lines. They are most probably associated with a strong Lyman-$`\alpha `$ line seen at the same velocity (see Fig. 12). Unlike other narrow systems seen over this wavelength range, the profile of the O vi absorption is well defined and less affected by blending. The equivalent widths of the C iv lines measured using a continuum defined locally, $`w`$ = 0.24 and 0.13 ร
, are consistent with a covering factor close to unity.
A Voigt-profile fit to the C iv$`\lambda `$1548 narrow line, with a continuum defined locally, gives $`b`$ = 11.$`\pm `$0.3 km s<sup>-1</sup>and $`N`$(C iv) = $`1.5\pm 0.1\times 10^{13}`$ cm<sup>-2</sup>. This gives an upper limit on the temperature of the gas, $`T9\times 10^4`$ K. The C iv$`\lambda `$1548 , Si iv$`\lambda `$1393 and O vi$`\lambda `$1037 profiles are very similar, suggesting that the velocity dispersions are of the same order. Higher order Lyman series lines are present exactly at the redshift defined by the C iv$`\lambda `$1548 line. The Ly-6 line is weak and not saturated. Voigt-profile fitting to this line gives $`N`$(H i) = 3.3$`\pm `$0.2 10<sup>15</sup> cm<sup>-2</sup> and $`b=20\pm 2`$ km s<sup>-1</sup>, which implies $`T2.4\times 10^4K`$ and suggests that ionization is dominated by photoionization.
#### 6.2.2 System at $`v9650`$km s<sup>-1</sup>
This system is defined by strong and broad C iv and Si iv lines. There is an absorption line at the expected position of Si iii$`\lambda `$1206 with very similar velocity profile which leaves little doubt on the identification (see Fig. 12). The fit to the Si iv doublet gives a covering factor of $``$0.25. For this value we estimate the column density of Si iv to be $`10^{14}`$ cm<sup>-2</sup>. It is interesting to note that the residual intensities of Si iii$`\lambda `$1206 and Si iv are similar. If both absorptions are produced in the same region this means that the column density of Si iv is about a factor of 3 larger than that of Si iii.
As the continuum fitting in the Si iv region is accurate, the estimate of the covering factor is reliable. The low-resolution data indicates that the Si iv absorption is not located on top of the Si iv emission line. Thus, a partial covering factor has to be explained in terms of one or two of the multiple images not being covered. However, the covering factor we derive here is larger than what is expected if the gas covers only image C and smaller than the expected value if it covers only image B. This most certainly means that the absorbing gas covers at least two of the images and that the optical depths along different sight lines are different. Absorptions due to the C iv lines are strong and both have residual intensity $``$0.20 (covering factor $`f0.80`$). Though this value is somewhat uncertain due to the non-uniqueness of the continuum fitting, it is larger than the value derived for Si iv. It has been observed already in some โmini-BALsโ that covering factors vary from one transition to the other (see e.g. the $`z_{\mathrm{abs}}`$= 2.207 system towards J2233-606, Petitjean & Srianand 1999). However, the interpretation of this in the present case is clearly that the ratio of the optical depths of Si iv and C iv along different sight-lines are different.
The Lyman-$`\alpha `$ line at about the same velocity is most certainly from an intervening system at $`z_{\mathrm{abs}}`$ $``$ 3.7584 associated with two weak C iv and Si iv narrow components (marked by dotted lines at $`9440`$ and $`9530`$ km s<sup>-1</sup> on Fig. 12). The determination of the H i column density in the broad system is difficult as the higher Lyman series lines are expected to be broad and shallow if the covering factor is 0.25. However we note that the residual flux at the expected positions of Ly-6 and Ly-7 are smaller than 0.25 suggesting that these lines, if at all present, are not heavily saturated. We obtain an upper limit on $`N`$(H i) of 4$`\times 10^{14}`$ cm<sup>-2</sup> for an assumed covering factor of 0.25. For a Mathews & Ferland (1987) ionizing spectrum, the ionization parameter required to produce the observed Si iv to Si iii column density ratio is log $`U2.00`$. This suggests that the absorbing gas is weakly ionized compared to the systems we have discussed above. The model needs \[Si/H\]$``$20\[Si/H\] in order to reproduce the column densities. Note that in this case the solution depends weakly on the assumed value of $`\alpha _{\mathrm{OX}}`$
### 6.3 Narrow components in the range $`9600<v<7000`$ km s<sup>-1</sup>
From the C iv profiles (see Fig. 9), it is apparent that this velocity range is dominated by absorptions from individual clouds superposed on a continuous absorption. It is difficult to disentangle both contributions. The narrow C iv systems have small H i optical depth. This is characteristic of systems associated with the quasar.
It is interesting to note on Fig. 9 that the residual flux of O vi$`\lambda `$1037 is very small and narrow components are barely visible. This indicates that the continuous gas-component is of higher ionization than the other components in the flow. Absorption due to Si iv is not detected over this velocity range. Thus, the systems are similar to the high-ionization systems discussed in Section 5.
### 6.4 Complex centered at $`v7000`$km s<sup>-1</sup>
As can be seen in Fig. 13, there are at least 10 distinct components (marked with dashed lines) in the C iv profile of this complex. The two strong Lyman-$`\alpha `$ lines at $`6800`$ and $`7100`$ km s<sup>-1</sup> are most probably due to intervening systems. The line at $`v`$ = $`6800`$ km s<sup>-1</sup> ($`z_{\mathrm{abs}}`$= 3.8018) is indeed associated with the five weak and narrow C iv and Si iv components with the largest velocities ($`>6850`$ km s<sup>-1</sup>).
The other C iv components have strong associated N v absorption with $`N`$(N v) $`>`$ $`N`$(Si iv), typical of associated systems. The non similarity of the N v$`\lambda `$1238 and N v$`\lambda `$1242 profiles indicates that there is some blending. As noted before, the expected position of Lyman-$`\alpha `$ from the two high-ionization systems at $`z_{\mathrm{abs}}`$$`3.90`$, discussed in Section 5, coincides with N v$`\lambda `$1242 in this system. Moreover, the N v$`\lambda `$1242 line of the system at $`v=8120`$ km s<sup>-1</sup>coincides with the blue wing of the N v$`\lambda `$1238 absorption. The strong absorption in the N v$`\lambda `$1242 profile at $`v=6990`$ km s<sup>-1</sup>is most certainly Lyman-$`\alpha `$ at $`z_{\mathrm{abs}}`$ = 3.904 with possible N v but no O vi associated absorption.
From the N v$`\lambda `$1238 profile it is clear that the gas producing the N v absorption covers more than 80% of the background source. Fitting of the unblended C iv and Si iv doublets results in covering factors in the range 0.8$``$0.9$`\pm 0.1`$. It is clear from Fig. 13 that the apparent optical depth of H i in the range $`7050<v<6900`$ km s<sup>-1</sup> is much smaller than the apparent optical depth of C iv and N v. This suggests that these systems are very similar to the systems discussed in Section 5.
### 6.5 The velocity range $`6000<v<4000`$ km s<sup>-1</sup>
The absorption profiles in the velocity range considered here is mainly due to broad lines (see Fig. 14). The component at $`v=5350`$ km s<sup>-1</sup> is defined by C iv$`\lambda `$1548 , N v and O vi absorptions. The width of the absorptions is consistent with the definition of โmini-BALโ. It can be seen on Fig. 14 that the optical depth increases dramatically from H i to O vi (note the weak but significant residual in O vi$`\lambda `$1037 ). This clearly suggests that $`N`$(O vi$`>>`$ $`N`$(C iv) and that the ionization of this system is very high.
The component at $`v=4700`$ km s<sup>-1</sup> is defined by C iv, Si iv, N v and O vi absorptions. The two lines of the C iv doublet are partially blended. The width of the lines is $`1500`$ km s<sup>-1</sup>. The Si iv$`\lambda `$1393 profile has a width of $`450`$ km s<sup>-1</sup>. The Si iv$`\lambda `$1402 line is blended with the C iv$`\lambda `$1548 from $`z_{\mathrm{abs}}`$= 3.3776. As the latter intervening system is itself heavily blended, unique profile decomposition is not possible and it is difficult to remove the C iv$`\lambda `$1548 contribution to the broad Si iv$`\lambda `$1402 profile. The important observation is that the residual intensities of O vi$`\lambda `$1031 , O vi$`\lambda `$1037 and N v$`\lambda `$1238 are larger than that derived for the component at $`v=5300`$ km s<sup>-1</sup> while it is the opposite for the C iv residual intensity. This probably indicates that this system is of lower ionization compared to the โmini-BALโ at $`z_{\mathrm{abs}}`$ = $``$5300 km s<sup>-1</sup>.
## 7 Line-locking
Several mechanisms have been proposed to explain the acceleration of BAL winds (see a review by de Kool 1997). One of the signatures of acceleration dominated by line radiation pressure is the presence of line-locking (e.g. Q 1303+308 in Foltz et al. 1987; NGC 5548 in Srianand 2000). We searched the spectrum of APM 08279+5255 for such signatures in the C iv wavelength range. We note that some of the velocity separations between the โmini-BALsโ are close to the Si iv and O vi doublet splittings. However as the individual lines are broad, the probability of such occurrence by chance is high. More striking is that most of the narrow C iv systems have a โcompanionโ system with a velocity difference corresponding to the velocity splitting of the Si iv, N v or O vi doublets with line centroids coinciding within 5 km s<sup>-1</sup>. The maximum number of matchings occurs at the velocity separation of the O vi doublet as expected for a high-ionization system. As an illustration we show in Fig. 15 the components with velocity separations equal to the O vi, N v and Si iv doublet separations. We believe that the data are indicative of the presence of โline-lockingโ.
## 8 Summary
### 8.1 Low-ionization systems
Two of the narrow absorption systems (at $`z_{\mathrm{abs}}`$ = 3.8931 and $`z_{\mathrm{abs}}`$ = 3.9135), show absorptions from singly ionized species (C ii, Si ii, Al ii) whereas absorptions from N v and O vi are weak or absent. Absorptions from excited fine structure levels of C ii (excitation energy corresponding to 156$`\mu `$m) and Si ii (excitation energy corresponding to 34$`\mu `$m) are detected at $`z_{\mathrm{abs}}`$ = 3.8931. The relative column densities are consistent with the shape of the IR spectrum of APM 08279+5255. Together with the low-ionization level of the system, this favors a picture where the cloud is closer to the IR source than to the UV source. This supports the idea that the extension of the IR source is larger than $``$200 pc. Alternatively, if the excitation is due to collisions, the electronic density must be larger than 100 cm<sup>-3</sup>. The dimensions of the cloud along the line-of-sight is of the order of 0.006 pc and $``$50/$`\sqrt{k}`$ times larger in the perpendicular direction in order for the covering factor to be $``$97% ($`k`$ is the amplification factor and could be as large as 100). Excited fine-structure lines are not detected in the $`z_{\mathrm{abs}}`$ = 3.9135 system. This suggests that the cloud is farther away from the QSO compared to the $`z_{\mathrm{abs}}`$ = 3.8931 cloud. Using photo-ionization models with and without attenuation by associated systems acting as a screen between the QSO and the gas, it is shown that abundances are $`ZZ_{}`$ and 0.01 $`Z_{}`$ at $`z_{\mathrm{abs}}`$ = 3.8931 and 3.9135 respectively. All this suggests that whereas the $`z_{\mathrm{abs}}`$= 3.8931 system is probably located within 200 pc from the QSO and ejected at a velocity larger than 1000 km s<sup>-1</sup>, the $`z_{\mathrm{abs}}`$= 3.9135 system must be farther away and part of the host-galaxy.
Aluminum could be over-abundant with respect to silicon and carbon at least by a factor of two and five at $`z_{\mathrm{abs}}`$ = 3.8931 and 3.9135 respectively. Such enrichment has already been observed by Ganguly et al. (1999) in a system with a velocity separation of 10000 km s<sup>-1</sup> with respect to Q1222+228. It is well established that in Galactic halo stars and field dwarfs the odd-Z elements (Na, Al, Cl, P, Mn and Co) have metallicities relative to iron smaller than solar by about a factor of 3 for iron metallicities in the range $`3.0[\mathrm{Fe}/\mathrm{H}]1.0`$ (Timmes et al. 1995). Conversely, $`\alpha `$chain elements (O, Ne, Mg, Si, S, Ar, Ca and Ti) generally have metallicities relative to iron larger than solar by about a factor of three for \[Fe/H\]$`1.0`$. The aluminum and silicon metallicities measured in damped Lyman-$`\alpha `$ systems follow these trends (Lu et al. 1997). In rapid star-formation models, massive star formation is favored and leads to high enrichment of aluminum and silicon with respect to carbon and nitrogen (Ferland et al. 1996). But \[Si/H\] is always larger than \[Al/H\]. Large \[Al/Fe\] ratios are observed in bright giant stars in mildly metal-poor globular clusters (Ivans 1999). Various correlations between different element abundances suggest that the abundance pattern is most likely due to very deep mixing and proton-capture nucleosynthesis (see Langer et al. 1997 for a review). However, although \[Al/Si\] is found larger than solar, the excess is quite small ($`<`$ 0.2 dex) and indeed much smaller than the factor 20 we observe in the $`z_{\mathrm{abs}}`$ = 3.9135 system.
Enrichment of aluminum with respect to silicon is found in the ejecta of classical novae (Andrea, Drechsel & Starrfield 1994; Gehrz et al. 1998). Therefore, the over enrichment of odd nucleus could be explained if the absorption is produced by individual novae shells (Shields 1996). However, novae shells invariably produce very high metal enrichment and high values of \[N/C\] compared to solar (Petitjean et al. 1990; Andrea et al. 1994), which conflict with the present observations.
### 8.2 High-ionization systems
The high-ionization associated systems at $`z_{\mathrm{abs}}`$ = 3.90โ3.917 (redshifts very close or even larger than the assumed intrinsic redshift, $`z_{\mathrm{em}}`$ = 3.91, see Fig. 8) have been shown to have metallicities larger than solar for a Mathews & Ferland ionization spectrum. At this redshift, the maximum life-time allowed by the $`\mathrm{\Omega }=1`$ cosmology is 0.8 Gyr, suggesting that rapid star formation occurs. We note that the relatively low N v to C iv and O vi column density ratios could be at odd with the standard rapid star-formation models discussed in the literature which predict larger values (Hamann & Ferland 1993; Matteucci & Padovani 1993). However it is known that BALQSOs are X-ray quiet with optical to X-ray spectral index $`\alpha _{\mathrm{OX}}1.9`$ (Green & Mathur, 1996) when the Mathews & Ferland spectrum corresponds to $`\alpha _{\mathrm{OX}}=1.40`$. If the cloud is ionized by a spectrum with $`\alpha _{\mathrm{OX}}1.9`$, observations are consistent with metallicities of the order of solar or slightly smaller.
The broad C iv absorption profile has a complex structure. It shows mini-BAL absorptions (of width $`1000`$ km s<sup>-1</sup>) and narrow components superposed on a continuous absorption of smaller optical depth. There is a tendency for mini-BALs to have different covering factors for different species. Again, this could be due to ionization structure of the BLR or due to ionization gradients in the flow. It is shown that the covering factors of a few of these absorbing clouds, which are well detached from the corresponding broad emission line profiles, are most probably due to the cloud not covering the three images of the lensed quasar. If true, this would put strong constraints on the dimensions of the clouds. HST spectroscopic observations of the three images are needed to confirm this findings. The continuous absorption is much stronger in O vi indicating that the diffuse component is highly ionized compared to the other components in the flow. This suggests that the absorption structures are due to density inhomogeneities within the flow.
The H i column density in the โmini-BALโ is small suggesting large metallicities as is commonly found in associated narrow as well as broad absorption systems. However, there is no absorption from Al iii and P v contrary to what is seen in some of the BAL systems (Junkkarinen et al. 1995). Indeed, Shields (1997) derived overabundance of aluminum compared to silicon, \[Al/Si\] = 6\[Al/Si\], in the BAL outflows of Q1101+091, Q1231+1325 and Q1331-0108, using photo-ionization models with a Mathews & Ferland (1987) spectrum and ionization parameters in the range $``$1.5 $`<`$ log $`U`$ $`<`$ 1.00. He also noted that there are indications for \[Al/Si\] being several times larger than solar in the BLR of some QSOs. The Al iii to Si iv optical depth ratio is less than 0.10 in the mini-BALs seen in the spectrum of APM 08279+5255. As the above ratio depends only weakly on the ionization parameter for log $`U0.50`$ (Hamann 1997), we can use a simple scaling of the analysis of Shields (1997) to derive that \[Al/Si\]$``$2 \[Al/Si\].
Finally we tentatively identify narrow components within the BAL-flow which have velocity separations very close (within 5 km s<sup>-1</sup>) to the O vi, N v and Si iv doublet splittings. This strongly suggests the existence of โline-lockingโ and points towards radiative acceleration as an important process in driving the outflow.
###### Acknowledgements.
We would like to thank the team headed by Sara L. Ellison to have made this beautiful data available for general public use. We gratefully acknowledge support from the Indo-French Centre for the Promotion of Advanced Research (Centre Franco-Indien pour la Promotion de la Recherche Avancรฉe) under contract No. 1710-1. |
warning/0003/astro-ph0003164.html | ar5iv | text | # Primordial Lithium Abundance as a Stringent Constraint on the Baryonic Content of the Universe
## 1 Introduction
The absolute abundances of <sup>4</sup>He, <sup>2</sup>D, and <sup>7</sup>Li synthesized in the first three minutes following the hot Big Bang provide the key to a determination of the universal baryon density via its relationship to the $`\eta `$ parameter. In order to refine observational estimates of the primordial levels of these light elements, previous attempts have been made to measure <sup>4</sup>He and <sup>2</sup>D abundances in extragalactic sites, where potential problems associated with correction for the effects of post-BBN chemical evolution could be minimized or avoided. However, the โlowโ value of <sup>4</sup>He/H reported for metal-poor extragalactic HII regions by Pagel et al. (1992) and Olive, Skillman, & Steigman (1997) stands in contrast to the โhighโ value of <sup>4</sup>He/H reported by Izotov & Thuan (1998). The โhighโ value for <sup>2</sup>D/H reported for high-redshift intergalactic HI clouds by Songaila et al. (1994), Carswell et al. (1994), and Rugers & Hogan (1996) appears at odds with the โlowโ values for <sup>2</sup>D/H from Tytler et al. (1996), Burles & Tytler (1998ab), and Burles et al. (1999). Detection of a high <sup>2</sup>D/H abundance (Webb et al. 1997; Tytler et al. 1999) for a gas cloud at rather low redshift ($`z`$0.7) has made the problem more complicated, because this is opposite to the expectation, based on one-zone models of chemical evolution, that the amount of <sup>2</sup>D is should decrease following BBN as the result of various destruction processes.
Given that two different values for both the primordial levels of <sup>4</sup>He and <sup>2</sup>D, leading to two distinguishable values of $`\eta `$, remain tenable at present, there is a pressing need for accurate estimate of the primordial level of <sup>7</sup>Li, which provides an independent constraint on $`\eta `$. Obtaining such a constraint has proven difficult for two reasons. First, the prediction of the primordial <sup>7</sup>Li abundance is a non-linear function of $`\eta `$, which formally permits the assignment of two values for $`\eta `$ at each level of measured primordial <sup>7</sup>Li. Second, <sup>7</sup>Li is several orders of magnitude less abundant than the other two light elements, so that high-precision observations of surface Li abundances are not possible except for nearby stars, which might have experienced the effects of chemical evolution in the Galaxy.
Despite these apparent difficulties, ever since the discovery of a roughly constant value of Li (<sup>6</sup>Li + <sup>7</sup>Li) abundance in a small sample of metal-poor dwarf stars by Spite & Spite (1982, the so-called โSpite Plateauโ), many groups (e.g., Ryan et al 1996; hereafter RBDT; Bonifacio & Molaro 1997) have attempted to better determine the appropriate primordial level of this element (see Spite 2000 for a recent review). The recent high precision (and homogeneously analyzed) data of Ryan, Norris, & Beers (1999, hereafter RNB) showed that the Spite plateau is in fact incredibly โthin,โ with an intrinsic star-to-star scatter in derived Li abundance $`\sigma <0.02`$ dex.
RNB also claimed the existence of a statistically significant slope of $`A`$(Li) versus \[Fe/H\] in the Spite Plateau at low metallicity (first detected by Thorburn 1994), apparently due to the influence of early Galactic chemical evolution. Ryan et al. (2000, hereafter RBOFN) have shown that the observed slope in the Spite plateau can be used to empirically constrain the total expected contribution from Galactic Cosmic Rays (GCRs) and supernovae (SNe). These authors showed that a simple one-zone model for chemical evolution produces a slope which is of similar magnitude to that which is observed. However, their evaluation is necessarily tied to the presently uncertain relationship between O and Fe abundances in the early Galaxy. Furthermore, their evaluation is based on a model which ignores the expected stochastic nature of early star formation which is likely to apply during the first 10<sup>7</sup> to 10<sup>9</sup> years of chemical evolution in the early Galaxy.
In this paper we employ the SN-induced chemical evolution model presented by Tsujimoto, Shigeyama, & Yoshii (1999; hereafter TSY) and Suzuki, Yoshii, & Kajino (1999; hereafter SYK), which has the great advantage of treating the production of heavy and light elements consistently in the inhomogeneous early Galaxy (ยง2). We recover the observed slope of $`A`$(Li) versus \[Fe/H\], and obtain an estimate of the primordial level of lithium (ยง3). We discuss the impact of our results in the context of the standard BBN model (ยง4).
## 2 The Model for Galactic Chemical Evolution
Recent observations of the most metal-deficient stars in the Galaxy reveal that the elemental abundance patterns in these stars seem to reflect the contributions from single SN events (Audouze & Silk 1995; McWilliam et al. 1995; Ryan, Norris, & Beers 1996; Norris, Beers, & Ryan 2000). These stars may have been formed in individual SN remnant (SNR) shells, at a time when the interstellar gas was not well-mixed throughout the halo. Based on this scenario, TSY presented a SN-induced chemical evolution model which successfully explains the observed large scatter of Eu abundances in very metal-poor stars. An important prediction of this model is that metallicity, especially \[Fe/H\], cannot be used as an age indicator at early epochs. Similar models of early Galactic chemical evolution have been studied by Argast et al. (2000). The clear implication is that when one considers elemental abundances of metal-poor stars a distribution of stellar abundances must be constructed in the context of an explicit model, rather than taking the evolution of elements assuming a well-mixed ISM, as is often done in simple one-zone models (e.g., Fig.4 in TSY).
SYK extended the SN-induced chemical evolution model for analysis of the evolution of the light elements produced by both primary and secondary processes involving GCRs, and demonstrated that this model also reproduces the observed trends of <sup>9</sup>Be and B (<sup>10</sup>B + <sup>11</sup>B) data. Thus, a SN-induced chemical evolution model appears suitable for a self-consistent investigation of the evolution of various elements in an inhomogeneous Galactic halo.
For the purpose of comparison with previous work, we here consider the evolution of Li (<sup>6</sup>Li + <sup>7</sup>Li) as predicted using the same model presented in SYK and TSY. Spallative and fusion reactions of GCRs produce both <sup>6</sup>Li and <sup>7</sup>Li, while the $`\nu `$-process of SNe produces <sup>7</sup>Li alone. The predictions of Woosley & Weaver (1995) for progenitors of different mass are used for the yields of the $`\nu `$-process, but the absolute values of these yields are reduced in order to match the observed <sup>11</sup>B/<sup>10</sup>B ratio (Vangioni-Flam et al. 1996; Vangioni-Flam et al. 1998). The transport of GCRs is calculated by the leaky box model (Meneguzzi et al. 1971), and the source spectrum of energetic particles of each element associated with SNe is taken from SYK:
$$q_i(E,t)\frac{E+E_0}{[E(E+2E_0)]^{\frac{\gamma +1}{2}}}_{\mathrm{max}(m_t,m_{SN,l})}^{m_u}๐m$$
$$\times \{M_{Z_i}(m)+Z_{i,\mathrm{g}}(t)f_{\mathrm{cr}}M_{\mathrm{sw}}(m,t)\}\frac{\varphi (m)}{A_im}\dot{M}_{}(t\tau (m)),$$
(1)
where $`E_0=931\mathrm{M}\mathrm{e}\mathrm{V}`$, $`M_{Z_i}(m)`$ is the mass of the $`i`$-th heavy element synthesized and ejected from a star with mass $`m`$, and $`\dot{M}_{}(t)`$ is the star formation rate at time $`t`$. GCRs are assumed to come from both swept-up material (accelerated in the SN shock front) and SN-ejecta; $`f_{\mathrm{cr}}`$ is a free parameter which represents the contributions to GCRs from these two sources. In order to reproduce the observed data of <sup>9</sup>Be (Boesgaard et al. 1999) and <sup>6</sup>Li (Smith, Lambert, & Nissen 1998), the value of the GCR spectral index is set to $`\gamma =2.7`$, and the GCR composition parameter is set to $`f_{\mathrm{cr}}=0.007`$ (Suzuki & Yoshii 2000). This parameter corresponds to the situation where 3.5% of the total GCRs arise directly from SNe ejecta. The absolute value of $`q_i(E,t)`$ is chosen so that the predicted <sup>9</sup>Be abundance agrees with log(Be/H)=$`13.5`$ at \[Fe/H\]$`=3.0`$ as observed by Boesgaard et al. (1999).
## 3 Results
Figure 1 presents a comparison of the observations of $`A`$(Li) from the compilation of RBDT, and the more recent observations of RNB, to the theoretical prediction of the frequency distribution of long-lived stars in the $`A`$(Li)โ\[Fe/H\] plane. In this comparison we have used our best estimate of the primordial Li abundance $`A(\mathrm{Li})_p=2.09`$ (justified below), and have selected only the data for stars with $`\mathrm{T}_{\mathrm{eff}}>6000\mathrm{K}`$ and \[Fe/H\]$`<1.5`$. Because the two data sets exhibit rather different random errors, we plot the data of RNB only in the top panel of Fig. 1, and the data of both RNB and RBDT in the bottom panel, with the same theoretical predictions overlayed.
The effects due to early Galactic chemical evolution of Li (which produces levels in excess of $`A(\mathrm{Li})_p`$) are seen already at \[Fe/H\]$`3`$; the predicted trend is quite consistent with the data. Our model confirms the existence of the slope of Li abundances toward the metal-rich side of the Spite Plateau, a result which be tested further once a larger sample of precision data are obtained for more metal-rich stars (Ryan et al. 2000). It should be emphasized that this increasing trend of Li abundance is naturally derived by use of the same model which fits the data of the other light elements such as <sup>6</sup>Li, <sup>9</sup>Be and B (<sup>10</sup>B+<sup>11</sup>B) in metal-poor halo stars. The average excess of total Li over the primordial <sup>7</sup>Li value amounts to 0.03 dex at \[Fe/H\] $`=2.5`$, 0.09 dex at \[Fe/H\] $`=2`$, and 0.24 dex at \[Fe/H\] $`=1.5`$. We note that the contribution of <sup>7</sup>Li production through the $`\nu `$-process is less than 10% of the total production rate of Li, which indicates that most of Li in our model of the early Galaxy is produced by GCR $`\alpha +\alpha `$ reactions.
Since our model gives a reasonable amount of Li produced by postโBBN processes, it can be used to predict the precise quantity of Li synthesized during the BBN era. We estimate the primordial Li abundance $`A(\mathrm{Li})_p`$ by constructing likelihood plots as a function of $`A(\mathrm{Li})_p`$. To establish this likelihood function, the data (including random errors) are compared with the predicted stellar distribution for alternative values of $`A(\mathrm{Li})_p`$ in the $`A`$(Li)โ\[Fe/H\] plane. The predicted probability distribution $`P(๐ฑ)`$ for stars located at $`๐ฑ`$=(\[Fe/H\], $`A(\mathrm{Li}))`$ is then calculated independently on each \[Fe/H\] grid. If we take $`g(๐ฑ๐ฑ_i,\sigma _{๐ฑ_i})`$ to represent the Gaussian distribution at $`๐ฑ`$ for the $`i`$-th star observed at $`๐ฑ_i`$ with errors of $`\sigma _{๐ฑ_i}`$, the likelihood $`L(A(\mathrm{Li})_p)`$ on each $`A(\mathrm{Li})_p`$ can be calculated as
$$\mathrm{log}\{L(A(\mathrm{Li})_p)\}=\mathrm{\Sigma }_i\mathrm{log}\{_{\mathrm{all}\mathrm{space}}P(๐ฑ)g(๐ฑ๐ฑ_i,\sigma _{๐ฑ_i})d^2๐ฑ\}.$$
(2)
The inset of the top panel of Fig. 1 shows the likelihood function established by using the data in RNB and RBDT. Based on our likelihood analysis, the 95% confidence region on the primordial Li abundance occurs for
$$A(\mathrm{Li})_p=2.09_{0.02}^{+0.01}.$$
(3)
We have examined the results using the RNB data and the RBDT data respectively, but there is little ($`<0.005`$ dex), if any, difference between them in the derived estimate of $`A(\mathrm{Li})_p`$.
So far we have not included any systematic errors in the determination of Li abundances which might arise from a variety of sources. Following RBOFN, we now explicitly include estimates of these errors, except for the still-controversial issue of stellar depletion. The estimated value and confidence interval for $`A(\mathrm{Li})_p`$ then becomes
$$A(\mathrm{Li})_p=2.07_{0.04}^{+0.16},$$
(4)
where we have considered the errors arising from the use of 1-D model atmospheres (+0.10 dex), the adopted convection treatment (+0.08 dex), non-LTE effects ($`0.02\pm 0.01`$ dex), and uncertainties in adopted $`gf`$-values ($`\pm 0.04`$ dex). The estimation of stellar abundances of Fe and Li also depends crucially on the adopted temperature scale. However, the \[Fe/H\] values we have taken from RNB and RBDT come from a variety of sources, with slightly different temperature scales, so a global correction is not practical. Note that a 100 K difference in temperature scales yields a 0.065 dex difference in derived lithium abundance (RNB). In particular, if one adopts the temperature scale of Alonso, Arribas, & Martinez-Roger (1996), which is on average 120K hotter than the scale in RNB, the derived Li abundances might be on the order of +0.08 dex higher. Except for the $`0.02`$ dex offset from non-LTE effects, we take quadratic sums for these positive and negative errors separately, and have derived $`A(\mathrm{Li})_p`$ in equation 4 as our estimate with the 95% confidence levels including systematic errors.
## 4 Summary and Discussion
The importance of our new estimates of the primordial level of <sup>7</sup>Li is that they are derived, for the first time, from a self-consistent model which explains both the <sup>6</sup>LiBeB observations and the small, but real, increasing trend of Li appearing at \[Fe/H\]$`>3`$. Figure 2 shows our preferred value of $`A(\mathrm{Li})_p=2.07_{0.04}^{+0.15}`$ with a horizontal line, and the allowed 95% confidence intervals by a box, along with previous estimates of <sup>4</sup>He and <sup>2</sup>D. The theoretical prediction of standard BBN (Thomas et al. 1994; Fiorentini et al. 1998) is superposed. It is interesting to note that our preferred value lies at the very bottom of the valley of the function of <sup>7</sup>Li abundance against $`\eta `$, which means that we can assign a single value for $`\eta `$, independent of the results from <sup>4</sup>He and <sup>2</sup>D. Our analysis indicates $`\eta (\mathrm{Li})=(1.73.9)\times 10^{10}`$, which corresponds to a universal baryonic density parameter<sup>1</sup><sup>1</sup>1If taking into accout uncertainties ($`1\sigma `$ errors) of nuclear reaction rates for the theoretical BBN calculation, the constraints become $`\eta (\mathrm{Li})=(1.44.5)\times 10^{10}`$ and $`\mathrm{\Omega }_bh^2=(0.531.7)\times 10^2`$. $`\mathrm{\Omega }_bh^2=(0.641.4)\times 10^2`$ with the Hubble constant expressed as $`h=H_0/100`$ km s<sup>-1</sup>Mpc<sup>-1</sup>.
The range of $`\eta (\mathrm{Li})`$ we obtain appears to agree best with that inferred from the reported โlow <sup>4</sup>Heโ + โhigh <sup>2</sup>Dโ measured from extragalactic sites, rather than the pair of reported โhigh <sup>4</sup>Heโ + โlow <sup>2</sup>Dโ. This range is also consistent with that inferred from the primordial value of <sup>4</sup>He obtained from recent observations of HII regions in the Magellanic Clouds (Peimbert & Peimbert 2000).
However, the possible effects of stellar depletion, which are not taken into account here, might still play a role. If one adopts the reported โhigh <sup>4</sup>Heโ and โlow <sup>2</sup>Dโ values as the correct ones with which to estimate $`\eta `$, then it follows that <sup>7</sup>Li has been destroyed through stellar evolution during the long lifetimes of metal-poor halo stars. From our Li constraint one might conclude that Li has been depleted by a factor of $`2.7_{1.4}^{+0.7}`$ as a result of stellar processing.
The required depletion factor, $``$ 2$``$3, is much larger than the prediction ($`<1.2`$; Deliyannis et al. 1990) of so-called standard stellar evolution models which only take into account classical surface convection as the origin of mixing in stellar interiors. In order to destroy lithium in deeper and hotter regions extra mixing processes are necessary. Among such processes, stellar rotation is thought to be one of the most effective candidates, but depletion factors $`1`$ dex inferred by rotation-induced mixing depend sensitively on the initial conditions of stellar rotation (Pinsonnealt et al. 1992). So, in general, this process predicts that one might expect to see a scatter about the Spite Plateau as large as the depletion factor itself, reflecting star-to-star differences in stellar rotation and other properties. Although recent models of rotation-induced mixing obtain more moderate depletion factors $`0.20.4`$ dex (Pinsonnealt et al. 1999), the expected scatter in the Spite Plateau would still seriously contradict the very small intrinsic scatter ($`\sigma <0.02`$ dex) observed by RNB.
All the above considerations indicate that, unless a novel process which is capable of significant and uniform depletion of stellar Li abundance is identified, our preferred value of $`A(\mathrm{Li})_p`$ should be taken as a stringent constraint on $`\eta `$ or $`\mathrm{\Omega }_b`$. The inferred value of the baryonic contribution to the density parameter is $`\mathrm{\Omega }_b0.010.02`$ ($`h=0.75`$).
This work has been supported in part by the Grant-in-Aid for the Center-of-Excellence research (07CE2002) of the Ministry of Education, Science, Sports, and Culture of Japan. TCB acknowledges partial support for this work from grant AST 95-29454 from the National Science Foundation.
<sup>5</sup><sup>5</sup>affiliationtext: The โhigh <sup>2</sup>Dโ was obtained through the observation of the cloud at $`z=2.80`$ in front of the quasar, Q0014+813. However, Burles et al. (1999) pointed out that the observed spectrum of this cloud could also be fitted with the model having D/H$`=`$0. |
warning/0003/hep-ph0003327.html | ar5iv | text | # Diquark composites in the color superconducting phase of two flavor dense QCD
## Abstract
We study the Bethe-Salpeter equations for spin zero diquark composites in the color superconducting phase of $`N_f=2`$ cold dense QCD. The explicit form of the spectrum of the diquarks, containing an infinite tower of narrow (at high density) resonances, is derived. It is argued that there are five pseudo-Nambu-Goldstone bosons (pseudoscalars) that remain almost massless at large chemical potential. These five pseudoscalars should play an important role in the infrared dynamics of $`N_f=2`$ dense QCD.
In his studies Dmitrij Vasilievich Volkov was always led by the beauty within the problems he considered. One of his passions was the theory of spontaneous symmetry breaking, in particular, the dynamics of Nambu-Goldstone (NG) particles (both bosons and fermions), to which he contributed a great deal to our present understanding . Therefore we think it is most appropriate to report in this volume a recent investigation of the dynamics of diquark composites (in particular, diquark NG bosons) in cold dense QCD with two fermion flavors.
Only a few years ago, not much was known about the properties of different phases in dense quark matter (see, however, Refs. ). The situation drastically changed after the ground breaking estimates of the color superconducting order parameter were obtained in Refs. . Within the framework of a phenomenological model, it was shown that the order parameter could be as large as 100 MeV. Afterwards, the same estimates were also obtained within the microscopic theory, quantum chromodynamics . The further progress in the field was mostly motivated by the hope that the color superconducting phase could be produced either in heavy ion experiment, or in the interior of neutron (or rather quark) stars.
Despite many advances in study of the color superconducting phase of dense quark matter, the detailed spectrum of the diquark bound states (mesons) is still poorly known. In fact, most of the studies deal with the NG bosons of the three flavor QCD. At best, the indirect methods of Refs. could probe the properties of the pseudo-NG bosons. It was argued in Ref. , however, that, because of long-range interactions mediated by the gluons of the magnetic type , the presence of an infinite tower of massive diquark states could be the key signature of the color superconducting phase of dense quark matter.
In this paper, we consider the problem of spin zero bound states in the two flavor color superconductor using the Bethe-Salpeter (BS) equations. We find that the spectrum contains five (nearly) massless states and an infinite tower of massive singlets with respect to the unbroken $`SU(2)_c`$ subgroup. Furthermore, in the hard dense loop improved ladder approximation, the following mass formula is derived for the singlets:
$$M_n^24|\mathrm{\Delta }|^2\left(1\frac{\alpha _s^2\kappa }{(2n+1)^4}\right),n=1,2,\mathrm{},$$
(1)
where $`\kappa `$ is a constant of order 1 (we find that $`\kappa 0.27`$), $`|\mathrm{\Delta }|`$ is the dynamical Majorana mass of quarks in the color superconducting phase, and $`\alpha _s`$ is the value of the running coupling constant related to the scale of the chemical potential $`\mu `$.
At large chemical potential, we also notice an approximate degeneracy between scalar and pseudoscalar channels. As a result of this parity doubling, the massive diquark states come in pairs. In addition, there also exist five massless scalars and five (nearly) massless pseudoscalars \[a doublet, an antidoublet and a singlet under $`SU(2)_c`$\]. While the scalars are removed from the spectrum of physical particles by the Higgs mechanism, the pseudoscalars remain in the spectrum, and they are the relevant degrees of freedom of the infrared dynamics. At high density, the massive and (nearly) massless states are narrow resonances.
In the case of two flavor dense QCD, the original gauge symmetry $`SU(3)_c`$ breaks down to the $`SU(2)_c`$ by Higgs mechanism. The flavor $`SU(2)_L\times SU(2)_R`$ group remains intact at the vacuum. The appropriate order parameter is an antitriplet in color and a singlet in flavor. Without loss of generality, we assume that the order parameter points in the third direction of the color space. In order to have a convenient description of the bound states at the true vacuum, we introduce the following Majorana spinors,
$`\mathrm{\Psi }_a^i=\psi _a^i+\epsilon _{3ab}\epsilon ^{ij}(\psi ^C)_j^b,`$ $`a=1,2,`$ (2)
$`\mathrm{\Phi }_a^i=\varphi _a^i\epsilon _{3ab}\epsilon ^{ij}(\varphi ^C)_j^b,`$ $`a=1,2,`$ (3)
made of the Weyl spinors of the first two colors,
$`\psi _a^i=๐ซ_+(\mathrm{\Psi }_D)_a^i,`$ $`(\psi ^C)_j^b=๐ซ_{}(\mathrm{\Psi }_D^C)_j^b,`$ (4)
$`\varphi _a^i=๐ซ_{}(\mathrm{\Psi }_D)_a^i,`$ $`(\varphi ^C)_j^b=๐ซ_+(\mathrm{\Psi }_D^C)_j^b.`$ (5)
Here $`i,j=1,2`$ are flavor indices, $`๐ซ_\pm =(1\pm \gamma ^5)/2`$ are the left- and right-handed projectors, $`\mathrm{\Psi }_D`$ is the Dirac spinor, and $`\mathrm{\Psi }_D^C=C\overline{\mathrm{\Psi }}_D^T`$ is its charge conjugate. Regarding the quark of the third color, we use the Weyl spinors, $`\psi ^i`$ and $`\varphi ^i`$, for left and right components, respectively (notice that the color index is omitted).
The BS wave functions of the bound diquark states in the channels of interest are given by
$`(2\pi )^4`$ $`\delta ^4(p_+p_{}P)๐_a^{(\stackrel{~}{b})}(p,P)=`$ (6)
$`=0|T\mathrm{\Psi }_a^i(p_+)\overline{\psi }_i(p_{})|P;\stackrel{~}{b}_L,`$
$`(2\pi )^4`$ $`\delta ^4(p_+p_{}P)๐_{(\stackrel{~}{a})}^b(p,P)=`$ (7)
$`=0|T\psi ^i(p_+)\overline{\mathrm{\Psi }}_i^b(p_{})|P;\stackrel{~}{a}_L,`$
$`(2\pi )^4`$ $`\delta ^4(p_+p_{}P)๐ผ(p,P)=`$ (8)
$`=0|T\mathrm{\Psi }_a^i(p_+)\overline{\mathrm{\Psi }}_i^a(p_{})|P_L,`$
$`(2\pi )^4`$ $`\delta ^4(p_+p_{}P)๐(p,P)=`$ (9)
$`=0|T\psi ^i(p_+)\overline{\psi }_i(p_{})|P_L,`$
where $`p=(p_++p_{})/2`$ and the quantities on the right hand side of these equations are defined as the Fourier transforms of the corresponding BS wave functions in the coordinate space. There are also the BS wave functions constructed out of the right handed fields $`\mathrm{\Phi }_a^i`$ and $`\varphi ^i`$. One might notice that there is another diquark channel, a triplet under $`SU(2)_c`$, that we do not consider here. The reason is that the repulsion dominates in such a channel, and no bound states are expected \[the triplet comes from the $`SU(3)_c`$ sextet\].
In order to derive the BS equations, we use the method developed in Ref. for the case of zero chemical potential. To this end, we need to know the quark propagators and the quark-gluon interactions.
By introducing the multicomponent spinor that combines the Majorana spinors of the first two colors and the Weyl spinors of the third color, $`(\mathrm{\Psi }_b^j,\psi ^j,\psi _j^C)^T`$, we find that the inverse propagator takes the following block-diagonal form:
$$G_p^1=\text{diag}(S_p^1\delta _a^b\delta _j^i,s_p^1\delta _j^i,\overline{s}_p^1\delta _i^j),$$
(10)
where, upon neglecting the wave functions renormalization of quarks ,
$`S_p^1`$ $`=`$ $`i\left(\overline{)}p+\mu \gamma ^0\gamma ^5+\mathrm{\Delta }_p๐ซ_{}+\stackrel{~}{\mathrm{\Delta }}_p๐ซ_+\right),`$ (11)
$`s_p^1`$ $`=`$ $`i\left(\overline{)}p+\mu \gamma ^0\right)๐ซ_+,`$ (12)
$`\overline{s}_p^1`$ $`=`$ $`i\left(\overline{)}p\mu \gamma ^0\right)๐ซ_{}.`$ (13)
Here the notation, $`\mathrm{\Delta }_p=\mathrm{\Delta }_p^+\mathrm{\Lambda }_p^++\mathrm{\Delta }_p^{}\mathrm{\Lambda }_p^{}`$, $`\stackrel{~}{\mathrm{\Delta }}_p=\gamma ^0\mathrm{\Delta }_p^{}\gamma ^0`$, and $`\mathrm{\Lambda }_p^\pm =(1\pm \stackrel{}{\alpha }\stackrel{}{p}/|p|)/2`$ are the same as in Ref. .
The bare vertex, $`\gamma ^{A\mu }`$, is also a $`3\times 3`$ matrix,
$$\gamma ^{A\mu }=\gamma ^\mu \left(\begin{array}{ccc}\overline{\gamma }_{11}^A\delta _j^i& \overline{\gamma }_{12}^A\delta _j^i& \overline{\gamma }_{13}^A\epsilon ^{ij}\\ \overline{\gamma }_{21}^A\delta _j^i& \overline{\gamma }_{22}^A\delta _j^i& \overline{\gamma }_{23}^A\epsilon ^{ij}\\ \overline{\gamma }_{31}^A\epsilon _{ij}& \overline{\gamma }_{32}^A\epsilon _{ij}& \overline{\gamma }_{33}^A\delta _i^j\end{array}\right),$$
(14)
with
$`\left(\overline{\gamma }_{11}^A\right)_a^b`$ $`=`$ $`T_a^{Ab}2\delta _8^AT_a^{8b}๐ซ_{},`$ (15)
$`\overline{\gamma }_{22}^A`$ $`=`$ $`T_3^{A3}๐ซ_+,`$ (16)
$`\overline{\gamma }_{33}^A`$ $`=`$ $`T_3^{A3}๐ซ_{},`$ (17)
$`\left(\overline{\gamma }_{12}^A\right)_a`$ $`=`$ $`T_a^{A3}๐ซ_+,`$ (18)
$`\left(\overline{\gamma }_{13}^A\right)_a`$ $`=`$ $`\epsilon _{3ac}T_3^{Ac}๐ซ_{},`$ (19)
$`\left(\overline{\gamma }_{21}^A\right)^b`$ $`=`$ $`T_3^{Ab}๐ซ_+,`$ (20)
$`\left(\overline{\gamma }_{31}^A\right)^b`$ $`=`$ $`T_c^{A3}\epsilon ^{3cb}๐ซ_{},`$ (21)
$`\left(\overline{\gamma }_{23}^A\right)_a`$ $`=`$ $`0,`$ (22)
$`\left(\overline{\gamma }_{32}^A\right)^b`$ $`=`$ $`0,`$ (23)
where $`T^A`$ are the $`SU(3)_c`$ generators in the fundamental representation. By making use of this vertex and the propagator in Eq. (10), it is straightforward to derive the BS equations in the (hard dense loop improved) ladder approximation. The details of the derivation, as well as the explicit form of equations are given elsewhere . Here we just note that the most transparent form of the equations appears for the amputated BS wave functions, defined by
$`\chi (p,P)=S^1(p+{\displaystyle \frac{P}{2}})๐(p,P)s^1(p{\displaystyle \frac{P}{2}}),`$ (24)
$`\lambda (p,P)=s^1(p+{\displaystyle \frac{P}{2}})๐(p,P)S^1(p{\displaystyle \frac{P}{2}}),`$ (25)
$`\eta (p,P)=S^1(p+{\displaystyle \frac{P}{2}})๐ผ(p,P)S^1(p{\displaystyle \frac{P}{2}}),`$ (26)
$`\sigma (p,P)=s^1(p+{\displaystyle \frac{P}{2}})๐(p,P)s^1(p{\displaystyle \frac{P}{2}}).`$ (27)
In order to get a feeling of the problem at hand, let us briefly discuss the analysis of the BS equation for the $`\chi `$-doublet. In general, the BS wave function contains eight different Dirac structures . It is of great advantage to notice that only four of them survive in the center of mass frame, $`P=(M_b,\stackrel{}{0})`$,
$$\chi _a^{(\stackrel{~}{b})}(p,0)=\delta _a^{\stackrel{~}{b}}\widehat{\chi }(p),$$
(28)
where
$`\widehat{\chi }(p)=[\chi _1^{}\mathrm{\Lambda }_p^++(p_0ฯต_p^{}+{\displaystyle \frac{M_b}{2}})\chi _2^{}\gamma ^0\mathrm{\Lambda }_p^+`$
$`+\chi _1^+\mathrm{\Lambda }_p^{}+(p_0+ฯต_p^++{\displaystyle \frac{M_b}{2}})\chi _2^+\gamma ^0\mathrm{\Lambda }_p^{}]๐ซ_+,`$ (29)
with $`ฯต_p^\pm =|\stackrel{}{p}|\pm \mu `$ \[the factors $`(p_0\pm ฯต_p^\pm +M_b/2)`$ are introduced here for convenience\]. This is the most general structure that is allowed by the space-time symmetries of the model.
Now, in the particular case of the NG bosons, $`M_b=0`$, we will show that the BS wave function is fixed by the Ward identities. Indeed, let us consider the following non-amputated vertex:
$$๐ช_{aj,\mu }^{A,i}(x,y)=0|Tj_\mu ^A(0)\mathrm{\Psi }_a^i(x)\overline{\psi }_j(y)|0,$$
(30)
where, for our purposes, it is sufficient to consider $`A=4,\mathrm{},8`$ (that correspond to the five broken generators). In the (hard dense loop improved) ladder approximation, the vertex satisfies the following Ward identity :
$$P^\mu ๐ช_{aj,\mu }^{A,i}(k+P,k)=iT_a^{A3}\delta _j^i\left[s_kS_{k+P}\right]๐ซ_{}.$$
(31)
As in the case of the BS wave functions, it is more convenient to deal with the corresponding amputated quantity,
$$\mathrm{\Gamma }_{aj,\mu }^{A,i}(k+P,k)=S_{k+P}^1๐ช_{aj,\mu }^{A,i}(k+P,k)s_k^1.$$
(32)
This latter satisfies the following identity:
$$P^\mu \mathrm{\Gamma }_{aj,\mu }^{A,i}(k+P,k)=iT_a^{A3}\delta _j^i\left[S_{k+P}^1s_k^1\right]๐ซ_+.$$
(33)
By making use of the explicit form of the quark propagators in Eqs. (11) and (12), we could check that the right hand side of Eq. (33) is non-zero in the limit $`P0`$. This is possible only if the vertex on the left hand side develops a pole as $`P0`$. After a simple calculation, we obtain
$$\mathrm{\Gamma }_{aj,\mu }^{A,i}(k+P,k)|_{P0}\frac{\stackrel{~}{P}^\mu }{P_\nu \stackrel{~}{P}^\nu }T_a^{A3}\delta _j^i\stackrel{~}{\mathrm{\Delta }}_k๐ซ_+,$$
(34)
where, we introduced $`\stackrel{~}{P}^\mu =(P_0,c_\chi ^2\stackrel{}{P})`$ with $`c_\chi `$ being the velocity of the NG boson in the $`\chi `$-doublet channel.
By making use of the definition in Eqs. (6) and (24), it is also not difficult to show that the pole contribution to the vertex function (34) is directly related to the BS wave function. By omitting the details,
$$\chi _a^{(\stackrel{~}{a})}(p,0)\delta _a^{(\stackrel{~}{a})}\chi (p,0)=\delta _a^{(\stackrel{~}{a})}\frac{\stackrel{~}{\mathrm{\Delta }}_p}{F^{(\chi )}}๐ซ_+,$$
(35)
where $`F^{(\chi )}`$ is the decay constant of the corresponding doublet whose formal definition is given by
$$0|\underset{A=4}{\overset{7}{}}T_a^{A3}j_\mu ^A(0)|P,\stackrel{~}{b}_L=i\delta _a^{\stackrel{~}{b}}\stackrel{~}{P}_\mu F^{(\chi )}.$$
(36)
By comparing the Dirac structures in Eqs. (29) and (35), we see that no components of the $`\chi _2^\pm `$ type appear in Eq. (35) which follows from the Ward identities. It was rewarding to establish that, in this approximation, the structure of the BS wave function required by the Ward identity is indeed a solution to the BS equation for the $`\chi `$-doublet. A similar situation takes place for the $`\eta `$-singlet .
Now let us discuss the fate of the massless states that we obtain. Altogether, there are five scalars and five pseudoscalars (a doublet, an antidoublet and a singlet). Because of the Higgs mechanism, the scalars are removed from the spectrum. Nevertheless, these scalar bound states exist in the theory as โghostsโ , and one cannot get rid of them completely, unless a unitary gauge is found. In fact, these ghosts play a very important role in getting rid of unphysical poles from the on-shell scattering amplitudes .
As for the pseudoscalars, they remain in the spectrum as pseudo-NG bosons. In the (hard dense loop improved) ladder approximation, they look like NG bosons because the left and right sectors of quarks decouple. One could think of this as an effective enlargement of the original color symmetry from $`SU(3)_c`$ to an approximate $`SU(3)_{c,L}\times SU(3)_{c,R}`$. Then, since the approximate symmetry of the ground state is $`SU(2)_{c,L}\times SU(2)_{c,R}`$, five scalar NG bosons (which are removed by the Higgs mechanism) and five pseudoscalar NG bosons (which remain in the spectrum) should appear. Of course, in the full theory, the pseudoscalars are only pseudo-NG bosons. Indeed, they should get non-zero masses due to higher orders corrections that are beyond the improved ladder approximation . Since the theory is weakly coupled at large chemical potential, it is natural to expect that the masses of the pseudo-NG bosons are small compared to the value of the dynamical quark mass.
We conclude our discussion of the massless diquarks by emphasizing that the low-energy dynamics of the two flavor QCD is dominated by massless quarks of the third color (which might eventually get a small mass too if another (non-scalar) condensate is generated ) and by the five pseudoscalars that remain almost massless in the dense quark matter. Of course, the gluons (glueballs) of the unbroken $`SU(2)_c`$ may also be of some relevance but we do not study this question here.
Now, let us consider massive diquarks. The structure of the BS equations becomes even more complicated in this case. In addition, one does not have a rigorous argument to neglect the component functions like $`\chi _2^\pm `$ in Eq. (29). In spite of this, we argue that all the approximations made before might still be reliable. Indeed, from the experience of solving the gap equation (which coincides with the BS equation for the massless states), we know that the most important region of momenta in the integral equation is $`|\mathrm{\Delta }|p\mu `$. In this region, the kernel of the BS equations for massive states, $`M_b<|\mathrm{\Delta }|`$, is almost the same. The deviations appear only in the infrared region where $`p<|\mathrm{\Delta }|`$.
Therefore, in our analysis of the BS equations for massive states, we closely follow the approximation used for the massless diquarks. By assuming that the component functions depend only on the time component of the momentum (compare with the analysis of the gap equation in Refs. ), we arrive at the following equation for the BS wave function of the massive singlets:
$$\eta _1^{}(p)=\frac{\alpha _s}{4\pi }_0^\mathrm{\Lambda }๐qK^{(\eta )}(q)\eta _1^{}(q)\mathrm{ln}\frac{\mathrm{\Lambda }}{|qp|},$$
(37)
where $`\mathrm{\Lambda }=(4\pi )^{3/2}\mu /\alpha _s^{5/2}`$, and the kernel reads
$$K^{(\eta )}(q)=\frac{\sqrt{q^2+|\mathrm{\Delta }|^2}}{q^2+|\mathrm{\Delta }|^2\left(M_\eta /2\right)^2},$$
(38)
\[$`\eta _1^{}(p)`$ is a scalar function that appears in the decomposition of the BS wave function $`\eta `$ over the Dirac matrices (compare with Eq. (29)\]. At this point it is appropriate to emphasize that the Meissner effect plays an important role in the analysis of the massive bound states. Indeed, our analysis shows that these massive states are quasiclassical in nature, i.e., their binding energy is small compared to the value of the gap \[see Eq. (1)\]. As a result, only the long range interaction mediated by the unscreened gluons of the unbroken $`SU(2)_c`$ is strong enough to produce these diquark states. We took this into account in Eq. (37). For completeness, we mention that the (nearly) massless (pseudo-) NG bosons are tightly bound, and the Meissner effect is not so important for their binding dynamics.
By approximating the kernel (38) in each of the following three regions: $`0<q<\sqrt{|\mathrm{\Delta }|^2\left(M_\eta /2\right)^2}`$, $`\sqrt{|\mathrm{\Delta }|^2\left(M_\eta /2\right)^2}<q<|\mathrm{\Delta }|`$ and $`|\mathrm{\Delta }|<q<\mathrm{\Lambda }`$, we could solve the BS equation (37) analytically. Then, by matching the logarithmic derivatives of the separate solutions, we obtain the spectrum of the massive diquarks. By omitting the details, it is presented in Eq. (1).
We would like to emphasize that for large $`\mu `$ the hard dense loop improved ladder approximation is reliable for the description of those bound states. The point is that a) the region of momenta primarly responsible for the formation of these composites is $`E_{bind}<q\mu `$, where the binding energy $`E_{bind}\alpha _s^2\mathrm{\Delta }`$, and b) $`E_{bind}\mathrm{}`$ as $`\mu \mathrm{}`$ . Therefore the vacuum effects are higher order ones in $`\alpha _s`$ in that region. Because of that, the hard dense loop improved ladder approximation, in which the contribution of the vacuum effects to the running of the coupling constant is neglected and only the running due to the polarization effects provided by the quark matter (non-zero $`\mu `$) is taken into account, is justifiable for large $`\mu `$.
Now, let us consider the case of massive diquarks in the doublet channel. As is easy to check, the binding interaction in this channel is exclusively due to the five gluons affected by the Meissner effect. The approximate BS equation looks similar to Eq. (37), but with a different kernel and $`|qp|`$ in the logarithm replaced by $`|qp|+c\mathrm{\Delta }`$ where $`c=O(1)`$ is a constant . At high density when the coupling constant is weak, this equation does not allow a non-trivial solution for $`M0`$. From the physical viewpoint, this indicates that the heavy gluons, with $`M_{gl}(\alpha _s\mu ^2\mathrm{\Delta })^{1/3}\mathrm{\Delta }`$, cannot provide a sufficiently strong attraction to form massive radial excitations of the NG and pseudo-NG bosons.
At the end, let us note that the massive diquark states may truly be just resonances in the full theory, since they could decay into the pseudo-NG bosons and/or gluons (glueballs) of the unbroken $`SU(2)_c`$. At high density, however, both the running coupling $`\alpha _s(\mu )`$ and the effective Yukawa coupling $`g_Y=|\mathrm{\Delta }|/F|\mathrm{\Delta }|/\mu `$ ) are small, and, therefore, these massive resonances are narrow.
In conclusion, in this paper we studied the problem of diquark bound states in the color superconducting phase of $`N_f=2`$ dense QCD. While the scalar NG bosons are ghosts in the theory, the pseudoscalar pseudo-NG bosons are physical particles that should play an important role in the infrared. We also obtained the spectrum of the massive narrow diquark resonances, whose existence would be a clear signature of the unscreened long range forces in dense QCD.
Acknowledgments. V.A.M. thanks V.P. Gusynin for useful discussions. The work of V.A.M. and I.A.S. was partly supported by the Grant-in-Aid of Japan Society for the Promotion of Science No. 11695030. The work of I.A.S. was supported by the U.S. Department of Energy Grants No. DE-FG02-84ER40153 and No. DE-FG02-87ER40328. The work of L.C.R.W. was supported by the U.S. Department of Energy Grant No. DE-FG02-84ER40153. |
warning/0003/astro-ph0003378.html | ar5iv | text | # Discovery of the optical counterpart and early optical observations of GRB9907121footnote 11footnote 1Based on observations collected at SAAO, Sutherland; ESO, Paranal and La Silla (ESO Programs 63.O-0618 and 63-O-0567); and AAT, Australia.
## 1. Introduction
The past two years have witnessed tremendous progress in our understanding of Gamma-Ray Burst (GRB) phenomena. Thanks to the ability of Beppo-Sax to provide arcminute sized error boxes for the bursts, the first optical counterpart was detected in February 1997 for GRB970228 (van Paradijs et al. 1997; Costa et al. 1997). The HST observations of this GRB provided the first clear indication that the GRB is associated with an external galaxy and is unrelated to its nuclear activity (Sahu et al. 1997a). Shortly thereafter, the redshift measurement for GRB 970508 proved beyond doubt that GRBs are extragalactic in nature (Metzger et al. 1997; Djorgovski et al. 1997). More than a dozen GRB optical counterparts have been detected since then, with redshifts as high as 3.42 in the case of GRB971214 (Kulkarni et al. 1998).
The extensive observations of GRB afterglows in X-ray, optical and radio wavelengths have been shown to be consistent with cosmological fire-ball models (e.g. Paczyลski & Rhoads 1993; Mรฉszรกros & Rees 1997; Wijers et al. 1997). However, the exact cause of the GRBs has remained elusive, and progress in determining the actual mechanism causing the bursts has been slow. The two leading models for GRBs involve the collapse of a massive star (e.g. Woosley 1997; Paczyลski 1998), and the merging of a neutron star with either another neutron star or a black hole (e.g. Eichler et al. 1989; Narayan, Paczyลski, & Piran 1992; Sahu et al. 1997b). The โisotropic equivalent energyโ of some GRBs is as high as 10<sup>54</sup> ergs, which exceeds the energy equivalent of the total mass involved in the latter model, and hence is thought to favor the former. However, the beaming factor of the emission of the GRB in different wavebands can be high (in both models), so the energetics alone may not be conclusive in favoring one of the models over the other. Since the lifetime of a very massive star is of the order of a million years or shorter, the GRBs are expected to be within or close to star-forming regions in the massive-star collapse model. On the other hand, since the kick-velocity of a newly formed neutron star is of the order of 200 $`\text{km\hspace{0.17em}s}^1`$, and since the neutron star, on average, is already about 10<sup>8</sup> years old at the time of the burst, GRBs are generally expected to be far from star-forming regions in the merging neutron-star model (Bloom, Sigurdsson & Pols 1999). Thus the location of the GRB with respect to the star-forming regions in the host galaxy could be a distinguishing feature which can help in settling the question of the cause of the GRBs. However, if the presence of dense interstellar material is prerequisite for the onset of the optical afterglow, then OTs would be observed only in star-forming regions regardless of whether GRBs result from neutron-star mergers or from the collapse of massive stars. In any case, if some OTs are found far away from star-forming regions, neutron-star merger model would be favored. If all OTs are found in star-forming regions, the situation is less clear, and depending on the model, one may need to investigate other aspects to find distinguishing features. For example, one consequence of the massive-star collapse model is that the GRB should be accompanied by an underlying supernova, whose brightness variation is distinct from the power-law decay behavior of the OT and may be detectable.
There is a clear bimodality in the observed burst durations: long bursts have timescales of about 10 to 200 seconds, and short bursts have timescales of 0.1 to 1 second (Kouveliotou et al. 1995). All the afterglows that have been discovered so far belong to the subgroup with long bursts (Fishman 1999). All the OTs for which the host galaxies have been observed โ although this sample is small โ are found to be in galaxies whose spectral indicators suggest star-forming activity (e.g. Fruchter et al. 1999). So there is an interesting possibility that the long-duration GRBs are caused by the collapse of rapidly-rotating massive stars, and hence they occur in star-forming regions. The short-duration bursts may be of a different origin, and may be caused by the coalescence of two neutron stars. The detection of optical counterparts of short duration bursts, and their observations would be important to understand this scenario better.
GRB990712, being at $`z=0.434`$ (Galama et al. 1999a; Hjorth et al. 2000), is one of the closest GRBs observed so far and hence provides a good opportunity to determine the possible presence of an underlying supernova, the location of the GRB within the galaxy, and the luminosity of the galaxy in different wavebands. In this paper we present the optical observations of GRB990712 leading to the discovery of its optical counterpart, and subsequent optical imaging observations taken in different filters during the first $``$35 days after the burst.
## 2. Observations
GRB990712 was simultaneously detected on 1999 July 12.69655 with the Gamma-Ray Burst Monitor and the WFC unit 2 onboard the BeppoSAX satellite. The burst lasted for about 30 seconds in both the $`\gamma `$-ray (40 - 700 keV) and X-ray (2-26 keV) energy ranges and had a double-peaked structure. While its intensity in $`\gamma `$-rays is moderate, it exhibited the strongest X-ray prompt emission observed to date (Heise et al. 1999).
The initial Beppo-SAX position, with an error circle of 5 arcmin radius, was revised to one with an error radius of only 2 arcmin. Our first observations were made at this latter position with the SAAO 1m telescope at Sutherland, South Africa on July 12.87 UT, about 4.16 hours after the burst during time generally dedicated to PLANET microlensing observations (Albrow et al. 1998). The optical image was taken through an R filter, with an integration time of 900 sec. The detector was a $`1024\times 1024`$ SITe CCD with an image scale of 0.309 arcsec per pixel and a total field of view of $`5.3\times 5.3`$ square arcmin. Comparison of the image with the Digitized Sky Survey (DSS) image for the same field showed the presence of a new source well above the sky background, that was absent in the DSS image. This new source had a brightness of R = 19.4 $`\pm `$ 0.1 mag, which is about 2 magnitudes brighter than the limiting magnitude of the DSS image, indicating that the new source was the optical counterpart of the GRB (Bakos et al. 1999a,b). Fig. 1 shows the discovery image, along with the DSS image taken of the same region. Also marked in the figure are the reference stars used for photometric calibration of the OT. A spectrum of this source obtained a few hours later revealed emission and absorption lines which were used to derive redshift of 0.434 for this source (Galama et al. 1999a). Subsequent observations by various groups showed the decaying nature of this source further confirming the identification of the OT with the GRB.
We continued the observations with a multi-wavelength optical follow-up campaign using the telescopes at SAAO, ESO and AAO. A log of all the observations used in this analysis along with the derived magnitudes of the OT are listed in Table 1. The measurements include the acquisition images taken for the spectroscopic observations with the ESO VLT shortly after the discovery of the OT (Vreeswijk et al. 1999, in prep.).
### 2.1. Astrometry and Photometry of the Optical Counterpart
An astrometric solution of the field was carried out using 10 reference stars in the image which are marked in Fig. 1. Their coordinates, as taken from the USNO 2.0 catalog which uses the Hipparcos frame of reference, are listed in Table 2. The pixel centroids of the reference stars were determined using two-dimensional Gaussian fits. These centroids were combined with the USNO coordinates to determine an astrometric solution of the field. The resulting position of the optical counterpart is RA(2000) = 22<sup>h</sup> 31<sup>m</sup> 53.$`{}_{}{}^{s}061\pm 0^s.011`$, Dec(2000)= $`73^d24^{}28.^{\prime \prime }58\pm \mathrm{\hspace{0.17em}0}^{\prime \prime }.05`$.
Since sky conditions deteriorated due to clouds just after the discovery observations of the OT, no standard star observations could be taken on that night. The images were calibrated through observations of the standard stars F203 and F209 (Menzies et al. 1989) taken on 13th July at SAAO. All the photometric measurements were carried out using the IRAF aperture photometry task PHOT. The photometry of a few secondary standards in the field (Fig. 1) was carried out using the standard star observations, and the photometry of all GRB observations was then performed via these secondary standards. Since only two standard stars were observed, the extinction correction due to differential airmass between the GRB and the standard star observations were not applied in the preliminary analysis (Bakos et al. 1999a,b). Instead, the photometric measurements taken with the adjacent 50cm telescope were used to get the extinction coefficients, which resulted in a correction of โ0.075 mag in V, โ0.06 mag in R, and โ0.02 mag in I. (This is consistent with the discrepancy pointed out by Kemp & Halpern 1999). Our final magnitudes of the secondary photometric standards are listed in Table 2, and Table 1 lists the derived magnitudes of the OT in different bands at different epochs including one extra measurement reported by Kemp and Halpern (1999).
## 3. Analysis
### 3.1. The Light Curve in Different Wavebands
All the photometric measurements listed in Table 1 are also shown in Fig. 2. The observed light curves in different bands clearly show a continuously-changing slope suggesting a additional component to the power-law decay of the OT as first noted by Hjorth et al. (1999a). In an approach similar to the one followed by Hjorth et al. (1999b), we have fitted this additional component in two ways: (i) with an underlying galaxy of constant brightness, and (ii) an underlying galaxy plus a supernova similar to GRB980425, appropriately scaled to the redshift of GRB990712.
First, a power law decline of the form $`f(t)t^\alpha `$ for the OT and a constant contribution from the background galaxy was used to fit the observed light curves simultaneously in different bands, with the same slope for all the bands. The resultant slope is $`0.97`$, and the total $`\chi ^2`$ is 47 for 39 d.o.f. (Relaxing the condition of the same slope in different bands does not alter the fit parameters significantly). The total $`\chi ^2`$ is, however, fairly flat (between 44 and 52) for any value of $`\alpha `$ between 0.95 and 1.02 beyond which the total $`\chi ^2`$ increases rapidly. Fig. 2 shows the best fit to the light curves. The derived magnitude of the underlying galaxy is not very sensitive to the exact value of $`\alpha `$ and are found to be $`\mathrm{V}=22.3\pm 0.05`$, $`\mathrm{R}=21.75\pm 0.05`$ and $`\mathrm{I}=21.35\pm 0.05`$ where the uncertainties represent the variation of the derived magnitudes as a result of changing $`\alpha `$ between 0.95 and 1.02.
One possible consequence of the massive-star model for the GRB is that one should observe an underlying supernova in the lightcurve of an OT. The physical connection of a GRB with a supernova (SN), first suggested by the discovery of a peculiar type Ic SN in the error box of GRB 980425 (Galama et al. 1998; Iwamoto et al. 1998), has been strengthened further by recent observations of other GRBs. Castro-Tirado and Gorosabel (1999) suggested that the light curve of GRB980326 resembled that of a SN, and indeed Bloom et al. (1999) showed that the late time light curve of GRB 980326 can be explained by an underlying SN1998bw type SN at a redshift of around unity. For the afterglow of GRB 970228, Reichart (1999) and Galama et al. (1999b) find that a power-law decay plus SN1998bw light curve redshifted to the distance of the burst, $`z=0.695`$ (Djorgovski et al. 1999), fits the observed light curve very well. In light of these new findings, we have also fitted the observed light curve of GRB 990712 assuming the presence of an underlying supernova.
In order to determine the contribution of the underlying supernova, we first calculate the expected V, R, I magnitudes of SN1998bw, placing it at the redshift of GRB 990712, $`z=0.434`$. This includes wavelength shifting and time profile stretching (both by a factor of 1+z), and rescaling the magnitudes for a distance corresponding to $`z=0.434`$, assuming $`\mathrm{\Omega }_0`$ = 0.2. To account for the wavelength shift correctly, we interpolate the redshifted UBVRI broadband flux spectrum of the SN for each bin in time with a spline fit, obtain the V, R, and I fluxes at their effective wavelengths, and convert these back to obtain magnitudes in the observerโs frame (Fukugita et al. 1995). The resultant SN light curves in different bands are shown at the bottom of Fig. 3. In order to fit the GRB light curve, the SN flux is subtracted from the observed GRB flux, and the residual fluxes are assumed to be due to the OT and the underlying galaxy. The procedure outlined earlier is then used to fit the light curves. The $`\chi ^2`$ in this case is 52 for 42 d.o.f., which is clearly higher than the model without the supernova. The resulting slope of the decay is $`0.96`$ which is quite similar but the magnitudes of the underlying galaxy are fainter: $`\mathrm{V}=22.7\pm 0.05`$, $`\mathrm{R}=22.25\pm 0.05`$ and $`\mathrm{I}=22.15\pm 0.05`$. The brightness of the OT of the GRB at a given time is also slightly lower, but the combined brightness of the OT and the underlying supernova (which is the quantity that can be measured if the OT can be resolved from the galaxy) is higher than the brightness of the OT in the absence of an underlying supernova.
It is important to note, however, that there is no evidence that the SNe possibly underlying the GRB afterglows have the same brightness or the same decay behavior. Our assumption that the underlying supernova is identical to SN1998bw is dictated by two reasons. First, this is the only SN associated with a GRB whose light curve has been monitored extensively. Second, the number of data points in our light curves does not allow us to vary the characteristics of the underlying SN light curve. So the fact that the $`\chi ^2`$ is higher in this case can be misleading since the true underlying supernova may be different from that of SN1998bw, making the resultant $`\chi ^2`$ higher for our simplified model. This chi-square analysis indicates that that either GRB990712 is not associated with an underlying supernova or the underlying supernova is much fainter than the supernova associated with GRB980425. However, the qualitative conclusion that the galaxy is expected to be fainter and the OT plus the SN is expected to be brighter in the presence of an underlying supernova is unlikely to change. Hence late time HST observations, in which the OT is well resolved so that the brightness of the OT and the galaxy can be estimated separately, or late time ground-based observations when the brightness of the OT is negligible, would greatly help in determining the presence of an underlying supernova.
In both the above scenarios (i.e. with and without an underlying supernova), the power-law index of the decay $`\alpha `$ is about $`1`$, making it one of the slowest decline rates of all the OTs observed so far. Since $`\alpha `$1 would lead to a divergence of the total energy integrated over time, the slope must steepen at later times. This has been observed for GRB990510 which declined with $`\alpha =0.82`$ at early times, later steepening with time (Harrison et al. 1999). Power-law decays are thought to arise from electrons shocked by the relativistic expansion of the debris into the ambient medium. In such a case, the information on the change of slope can be used to derive information on the cooling rate of the electrons in the post-shock region (see, e.g. Wijers et al. 1997; Sari, Piran, & Narayan 1998; Livio & Waxman 1999).
### 3.2. Spectral Energy Distribution of the OT
If the OT emission is due to the synchrotron radiation from the swept-up electrons in the post-shock region, then its energy distribution is expected to be of the general form $`f_\nu t^\alpha \nu ^\beta `$. In such a case, one expects a relationship between the power law index of the energy distribution of the electrons $`p`$, the spectral slope $`\beta `$, and the decay constant $`\alpha `$ (Sari et al. 1998). One must distinguish two cases: (i) both the peak frequency $`\nu _\mathrm{m}`$ and the cooling frequency $`\nu _\mathrm{c}`$ are below the optical/IR waveband, in which case $`p=(4\alpha +2)/3=1.97\pm 0.04`$ and $`\beta =p/2=0.99\pm 0.02`$, and (ii) $`\nu _\mathrm{m}`$ has passed the optical/IR waveband, but $`\nu _\mathrm{c}`$ has not yet done so, in which case $`p=(4\alpha +3)/3=2.31\pm 0.04`$ and $`\beta =(p1)/2=0.66\pm 0.02`$.
We can now compare the theoretically expected spectral slope ($`\beta `$) with the observed one. Although perhaps best determined by the spectroscopic measurements. our light curve analysis provides magnitudes of the OT in different bands, which can be directly used to derive $`\beta `$. We should note here that the brightness of the underlying galaxy and the possible contamination by an underlying supernova makes the determination of the magnitudes of the OT less reliable at later times, and our value of $`\alpha `$ is most likely dominated by the early part of the light curve when the OT was bright. Furthermore, $`\alpha `$ was kept fixed in $`all`$ the bands for the entire period of our observations, which implies that we cannot detect any spectral evolution of the OT. The $`synphot/calcphot`$ task in IRAF was used to determine the V$``$R and R$``$I values for a range of $`\beta `$ values, which were then compared with the colors derived from the light curve analysis to determine the spectral slopes. The V$``$R color implies a spectral slope of $``$1.8 $`\pm `$ 0.5, and the R$``$I color implies a spectral slope of $``$0.7 $`\pm `$ 0.1. Thus the slope derived from the V and R bands is closer to the case (i) mentioned above, and the slope derived from the R and I bands is closer to the case (ii). This is roughly consistent with the theoretical expectations and implies that, in the early part of the light curve, $`\nu _m`$ has passed the V and R and I bands, but $`\nu _c`$ has not passed the I band.
### 3.3. The Background Galaxy
The inferred magnitudes for an underlying galaxy suggest that the galaxy is relatively bright compared to the OT, and it may be possible to see the host galaxy directly in the images. The NTT images taken about 8 days after the burst had the best seeing ($``$ 1 arcsec) and the R images indeed show some hint of a slight extension. To further investigate this, all the images taken in B,V,R and I bands were co-added, giving rise to a single deep image of a total integration time of 40 min. From this combined image, we constructed a model point-spread-function (PSF) from a few isolated bright stars in the field using IRAF/DAOPHOT. This model PSF was used to subtract the point source contribution from the OT. The PSF-subtracted image shows some residual distribution elongated in the east-ward direction of the OT over a total of about 2 arcsec, which is probably the contribution of the underlying galaxy. Thus HST observations would show the detailed structure of the galaxy (Fruchter et al. in prep).
The derived magnitudes of the underlying galaxy were used to calculate its luminosity. The GRB is at a Galactic latitude and longitude of โ40 and 315 degrees, respectively. The extinction models of Burstein & Heiles (1982) and Schlegel et al. (1998) yield E(BโV) of 0.015 and 0.033, respectively, for the position of the GRB. This extinction is small and similar to SN1998bw, which is neglected in our analysis. To derive the luminosity of the host galaxy, we use a redshift of z=0.434 which, for H<sub>0</sub> = 70 $`\text{km\hspace{0.17em}s}^1`$and $`\mathrm{\Omega }`$=0.2, corresponds to a luminosity distance of 2160 Mpc, and a distance modulus of 41.67 mag. Applying the K-correction for z=0.434, the inferred luminosity of the galaxy is of the order of $`L_{}`$ (depending on the presence of an underlying supernova), where $`L_{}`$ corresponds to the luminosity of a typical galaxy at the โkneeโ of the observed galaxy luminosity function (see, e.g., Lilly et al. 1999). If we exclude SN1998bw (which is not only โnearbyโ ($`z=0.008`$), but for which the energy released in $`\gamma `$-rays is much smaller than any other GRB), then GRB990712 is the closest โcosmologicalโ GRB, and the apparent magnitude of its host galaxy is the brightest among all the GRB host galaxies observed so far. Compared to the host galaxy of GRB990713, which is the next brightest, the host galaxy of GRB990712 is about 0.25 magnitudes brighter in R if there is an underlying supernova, and more than a magnitude brighter if there is no underlying supernova.
We dedicate this paper to the memory of our colleague and friend Jan van Paradijs who passed away due to an illness on November 2, 1999, during the final stages of the preparation of this manuscript. Jan received several awards for his fundamental contributions in this field, including the 1998 Rossi prize of the American Astronomical Society. We will sorely miss his contagious enthusiasm, his brilliant insights, and his benevolent personality. This project is partly supported by the Chilean Fondecyt Project No. 3990024 and the Danish Natural Science Research Council (SNF). M. Dominik is supported by a Marie Curie Fellowship (ERBFMBICT972457). |
warning/0003/astro-ph0003235.html | ar5iv | text | # NONTHERMAL EMISSION FROM A SUPERNOVA REMNANT IN A MOLECULAR CLOUD
## 1. INTRODUCTION
Supernovae and their remnants have long been identified as likely sites of galactic cosmic ray acceleration. Evidence for electron acceleration first came from observations of radio synchrotron radiation. The observation of nonthermal X-ray emission from SN 1006 by ASCA has provided evidence for electron acceleration up to $`100`$ TeV (Koyama et al. 1995; Reynolds 1998). The proton cosmic ray component is more difficult to observe, but it has long been recognized that pion decays from collisions with interstellar gas could give an observable flux of $`0.11`$ GeV photons. Initial estimates of $`\gamma `$-ray emission from supernova remnants (SNRs), concentrating on the component resulting from pion decays, were made by Chevalier (1977), Blandford & Cowie (1982), and Drury, Aharonian, & Vรถlk (1994). When $`\gamma `$-ray emission was apparently detected from SNRs by CGRO, the $`\gamma `$-ray spectrum could not be fitted by a pure pion decay spectrum and some other component was needed (Esposito et al. 1996; Sturner et al. 1997; Gaisser, Protheroe, & Stanev 1998). In addition to the pion decays, the relevant processes are bremsstrahlung emission of relativistic electrons and inverse Compton emission. Gaisser et al. (1998) modeled these processes in detail in order to fit the observed $`\gamma `$-ray spectra of the remnants IC 443 and $`\gamma `$ Cygni. They assumed acceleration to a power law spectrum in the shock front and determined the spectral index, electron to proton ratio, and the upper energy cutoff.
The evolution of the relativistic component in a supernova remnant has been modeled by a number of groups. Sturner et al. (1997) assumed particle acceleration in the shock front with an $`E^2`$ energy spectrum and followed the evolution of the particle spectrum. Their SNR model was homogeneous: a factor 4 density jump at the shock front, constant density in the interior, and a constant magnetic field in the interior. Their shock acceleration model had a diffusion length parameter that was important for the maximum energy reached by the particles. They modeled the nonthermal emission from IC 443 as synchrotron emission in the radio and bremsstrahlung in $`\gamma `$-rays. de Jager & Mastichiadis (1997) dealt with the same processes in W44, as well as inverse Compton emission. They noted that the observed radio spectrum is flatter than would be expected from shock acceleration of newly injected particles and suggested that the particles originated from a pulsar in the supernova remnant. Recently, Ostrowski (1999) showed that second-order electron acceleration by the turbulent medium just after the shock could flatten the spectra to account for the radio observations of IC 443. Baring et al. (1999) presented calculations of the broad-band emission from nonlinear shock models of shell-type SNRs. They used Sedov adiabatic shock dynamics in a homogeneous medium and a Monte Carlo simulation of the particle acceleration taking into account the nonlinear shock structure. The set of models considered by Baring et al. (1999) covers the range of shock speeds 490 $`v_S`$ 4000 km s<sup>-1</sup> and ambient medium number densities 10$`{}_{}{}^{3}n`$ 1 cm<sup>-3</sup>.
Massive stars that are the likely progenitors of core collapse supernovae are expected to be spatially correlated with molecular clouds. The remnants that are likely to be $`\gamma `$-ray sources in CGRO observations (Esposito et al. 1996) also show evidence for interaction with molecular gas. Chevalier (1999) recently studied the evolution of supernova remnants in molecular clouds and concluded that many aspects of the multiwavelength observations could be understood in a model where the remnants evolve in the interclump medium of a molecular cloud, which has density of $`525`$ H atoms cm<sup>-3</sup>, and become radiative at radii $`6`$ pc. The compression in the radiative shell is limited by the magnetic field. Molecular emission occurs when the radiative shell collides with molecular clumps. IC 443 is the remnant with the best evidence for high energy emission and also shows clear evidence for interaction of the shock with a molecular cloud (e.g., Burton et al. 1990; van Dishoeck et al. 1993; Cesarsky et al. 1999).
Most of the current models of particle acceleration in SNRs have dealt with adiabatic remnants in a homogeneous medium (see, however, the model of Boulares & Cox 1988 for the Cygnus Loop and the model of Jones & Kang 1993). In contrast, our aims here are to examine in detail the nonthermal emission of radiative supernova remnants in molecular clouds using a kinetic model of electron injection, acceleration and propagation, including the inhomogeneous structure deduced from multiwavelength SNR observations. We use global MHD (magnetohydrodynamic) models of supernova remnants interacting with molecular clouds as described by Chevalier (1999), including the theory of radiative shock structure by Shull & McKee (1979) and the kinetic model of electron injection and acceleration by shocks from Bykov & Uvarov (1999). The model described here is relevant to the mixed-morphology SNRs (Rho & Petre 1998), if they are interacting with molecular clouds. These remnants comprise a sizable fraction of the Galactic SNR population.
The plan of our paper is as follows. In ยง 2, we treat the energy spectrum of energetic electrons as they pass through the regions associated with a radiative shock wave. We include the shock acceleration of particles from the thermal pool. In ยง 3, we present calculations of the nonthermal emission from the relativistic particles, including emission from shocks in clumps as well as emission from the radiative shock wave in the interclump region. Our model is specifically applied to the well-observed remnant IC 443, although we expect it to more generally apply to other supernova remnants in molecular clouds. We also examine the ionization by energetic particles and the energy involved in the relativistic particles. A discussion of results and future propects is in ยง 4. A detailed discussion of expected range of densities, shock velocities and magnetic fields determined from multiwavelength observations of IC 443, W44 and 3C391 was given in the paper by Chevalier (1999, and references therein).
## 2. EVOLUTION OF NONTHERMAL ELECTRONS
The structure of the flow expected in a radiative shock can be divided into the following regions (see, e.g., Draine & McKee 1993):
(I) preshock gas ionized and heated by UV (ultraviolet) emission and fast particles from the shock;
(II) shock transition region;
(III) postshock cooling layer;
(IV) cold shell of swept-up gas;
(V) hot low density gas interior to the shock.
Radiative shocks are subject to thermal instabilities (Chevalier & Imamura 1982; Bertschinger 1986) as well as dynamical instabilities (Vishniac 1983; Blondin et al. 1998). Although the effect of inhomogeneous magnetic fields on the stability of radiative shock flow is yet to be studied, MHD turbulence is expected in such a system. The transport and acceleration of nonthermal particles in the violent SNR environment are governed by the MHD flow and depend on the ionization structure and on the spectrum of MHD turbulence. Shocks with velocity $`v_{S7}`$ 1.1 generate sufficient ultraviolet UV radiation for full pre-ionization of hydrogen and helium (He<sup>+</sup>) in the preshock region, where the gas temperature is typically $`10^4`$ K (Shull & McKee 1979). Here, $`v_{S7}`$ is the shock velocity measured in units of $`100\mathrm{km}\mathrm{s}^1`$, and $`n_1`$ is the ion number density measured in units of 10 cm<sup>-3</sup>. Shocks with $`v_{S7}`$ 1.5 propagating in a medium with $`n_1`$ 1 photoionize gas ahead of the viscous jump to a depth $`N_i>\mathrm{\hspace{0.33em}3}\times 10^{17}`$ cm<sup>-2</sup>. The shock transition in such a case is collisionless and supercritical unless the preshock magnetic field is above $``$ 4$`\times `$10$`{}_{}{}^{5}n_{1}^{0.5}`$ G.
The geometry of the magnetic field is important for the efficiency of electron injection and acceleration. It has been argued (e.g., Levinson 1996; McClements et al. 1997) that quasi-perpendicular shocks are an efficient source of freshly injected electrons if their speed $`v_s\left(m_e/m_p\right)^{0.5}c`$; the injection efficiency drops sharply below that threshold. On the other hand, Bykov & Uvarov (1999) have shown that quasi-parallel MHD shocks of moderate Alfven Mach number $`M_a\left(m_e/m_p\right)^{0.5}`$ are suitable sites of electron injection and acceleration. The SNR shock waves in molecular clouds are expected to be in this regime of Mach number. Thus we shall discuss below the quasi-parallel portion of the shock (magnetic field inclinations $`\pi /4`$) propagating into a molecular cloud with $`v_{S7}1.5`$.
Nonresonant interactions of the electrons with fluctuations, generated by kinetic instabilities of the ions in the transition region of a quasi-parallel supercritical shock, play the main role in heating and preacceleration of the electrons. The electron heating and preacceleration (injection) occur in the collisionless shock front transition region on a scale $`\mathrm{\Delta }`$ of several hundred inertial lengths of the ions, $`l_i`$ (Bykov & Uvarov 1999). Here $`l_i=c/\omega _{pi}9\times 10^6n_1^{0.5}`$ cm and $`\omega _{pi}`$ is the ion plasma frequency. For a SNR shock in a molecular cloud, the scale $`\mathrm{\Delta }`$ 10<sup>9</sup> cm is much shorter then the scales of all the other regions.
To accelerate electrons injected in the shock transition region to relativistic energy $`E(p)`$, MHD turbulence should fill the acceleration region, which has a scale $`l_ak_m(p)/v_s`$. Here, $`k_m=\mathrm{max}\{k_i\}`$, where $`k_i(p)`$ is the diffusion coefficient of an electron of momentum $`p`$ and $`i=`$(I), (III) refers to the regions defined above. To provide the conditions for efficient transformation of the MHD flow power to the accelerated electrons, the MHD turbulent fluctuation spectrum in the vicinity of the shock should extend to scales $`lr_e(p)`$ (resonant scattering), where $`r_e(p)`$ is the gyroradius of an electron with momentum $`p`$. The MHD fluctuations responsible for the electron scattering are collisionless for particles with energies up to a GeV because the ion-neutral collision length is above 10<sup>14</sup> cm for typical parameters of the radiative shock described below (a keV electron has $`r_e`$ 10$`{}_{}{}^{8}[B/10^6\mathrm{G}]_{}^{1}`$ cm).
To calculate the spectra of nonthermal electrons in the regions i = I $``$ IV, we use a kinetic equation for the nearly-isotropic distribution function $`N_i(z,p,t)`$:
$`{\displaystyle \frac{}{t}}N_i+u_i(z){\displaystyle \frac{}{z}}N_i{\displaystyle \frac{p}{3}}{\displaystyle \frac{}{p}}N_i\left({\displaystyle \frac{}{z}}u_i\right)=`$
$`k_i(p){\displaystyle \frac{^2N_i(z,p)}{z^2}}+{\displaystyle \frac{1}{p^2}}{\displaystyle \frac{}{p}}p^2D_i(p){\displaystyle \frac{N_i}{p}}+{\displaystyle \frac{1}{p^2}}{\displaystyle \frac{}{p}}[p^2L_i(p)N_i].`$ (1)
The Fokker-Planck type equation (1) takes into account diffusion and advection (with bulk velocity $`u_i[z]`$) of an electron in phase space due to particle interaction with MHD waves and the large scale MHD flow (Bykov & Toptygin 1993). Here, $`L_i(p)`$ is the momentum loss rate of an electron due to Coulomb collisions in a partially ionized plasma (e.g., Ginzburg 1979). The momentum diffusion coefficient $`D(p)`$ is responsible for second order Fermi acceleration, and $`k_i(p)`$ is the fast particle spatial diffusion coefficient. For low energy electrons (i.e., $`EE_C`$), Coulomb and ionization losses are important in the regions (I $``$ V), except for the narrow shock transition region (II) where acceleration is fast enough to overcome losses and nonthermal electron injection occurs. The characteristic energy $`E_C`$ depends on the plasma density and ionization state, magnetic field, and fast particle diffusion coefficients $`k_i(p)`$ and $`D_I(p)`$. The corresponding momentum $`p_{ci}`$ can be estimated as the point where the last two terms in equation (1) are equal. We used the energy $`E_C`$ calculated simultaneously with the electron distribution function as a convenient parameter to distinguish between different possible models of MHD turbulence in the postshock cooling layer of the radiative shock structure (see below in this section).
The diffusion coefficients $`k_I(p)`$ and $`D_I(p)`$ depend on the spectrum of the collisionless MHD turbulence, which is poorly understood. A plausible approximation for the diffusion coefficients $`k_i(p)`$ was assumed here. For all of the regions we used the following parameterization:
$$k_i(p)=k_{i0}\{\begin{array}{ccc}1,\hfill & p_Tpp_{}\hfill & \\ vp^a/v_{}p_{}^a,\hfill & p_{}pp_{}\hfill & \\ cp^2/(p_{}^{2a}p_{}^av_{})\hfill & p_{}pp_m.\hfill & \end{array}$$
(2)
Here, $`p_T`$ is the momentum of the upstream thermal electrons, $`p_m`$ is the momentum corresponding to the upper cut-off energy $`E_m`$ (see equation ), and $`p_{}`$ and $`p_{}`$ are defined in the next paragraph. The standard relation for the momentum diffusion coefficient $`D_i(p)`$ = $`p^2w_i^2/9k_i(p)`$ (e.g., Berezinsky et al. 1990) was used. In the low energy regime $`pp_{}`$, the particle transport is dominated by large scale turbulent advection (Bykov & Toptygin 1993). The large scale turbulence is due to stochastic MHD plasma motions on scales $`\mathrm{\Lambda }`$ longer than the particle mean free path due to resonant scatterings. The sources of MHD turbulent motions are the shock wave instabilities mentioned at the beginning of this section. The large scale vortex rms-velocity $`w_i`$ typically is a fraction of bulk speed $`u_i`$. The spatial diffusion is energy independent below $`p_{}`$ and $`k_{i0}w_i\mathrm{\Lambda }`$. For $`p_{}pp_{}`$, electrons are scattered by resonant MHD waves presumably generated by the streaming instability of shock accelerated particles (e.g., Blandford & Eichler 1987; Jones & Ellison 1991). In this case $`w_i`$ is close to the Alfven velocity $`v_a`$ and $`k_i(p)`$ is the electron diffusion coefficient due to resonant wave-particle interactions. If the power spectrum of magnetic field fluctuations of Alfvenic turbulence for the resonant fluctuations is approximated as $`dB_k^2/dkk^\theta `$ then the power-law index $`a`$ (equation ) is related to the index $`\theta `$ as $`a=2\theta `$. An important particular case corresponding to $`a=1`$ can be written as $`k_i(p)=(1/3)\eta _ivr_e(p)`$. The parameter $`\eta _i`$ determines the scattering โstrengthโ and strong scattering (Bohm limit) corresponds to $`\eta _i`$ 1. The diffusion model with $`a=1`$ was successfully used for modeling the observed anomalous cosmic ray proton fluxes in the interplanetary medium (Ellison et al. 1999) and in the Monte Carlo simulations of nonlinear shocks in the shell type SNRs (e.g., Baring et al. 1999). The momentum $`p_{}`$ in our model can be estimated from the equation $`(1/3)\eta _ivr_e(p_{})=w_i\mathrm{\Lambda }`$.
MHD waves in the partially ionized plasma of the radiative shock flow are subject to ion-neutral damping (e.g., Vรถlk et al. 1981). We define the momentum $`p_{}`$ from the condition $`r_e(p_{})=\lambda _{in}`$, where $`\lambda _{in}`$ is determined from the relation $`\lambda _{in}=2\pi v_a/\mathrm{\Gamma }_{in}`$ and $`\mathrm{\Gamma }_{in}`$ is the Alfven wave damping rate due to ion-neutral collisions (Kulsrud & Cesarsky 1971). For $`pp_{}`$, electrons are scattered by nonresonant small scale waves.
The limits imposed on the diffusive shock acceleration of particles due to ion-neutral MHD wave damping have been discussed by Draine & McKee (1993) and Drury et al. (1996). They considered the spectra of Alfven waves driven by the instabilities of shock accelerated ions in a partially ionized plasma. Drury et al. (1996) obtained the upper cut-off energy of a proton due to ion-neutral wave damping
$$E_mv_{S7}^3T_4^{0.4}n_n^1n_i^{0.5}๐ซ_1^{CR}\mathrm{GeV},$$
(3)
where $`n_n`$ is the neutral particle density, $`n_i`$ is the ion density (both are measured in $`\mathrm{cm}^3`$), and $`๐ซ_1^{CR}`$ is the total particle pressure normalized to 10% of the shock ram pressure. The ionization structure of the preshock region is inhomogeneous so the ion number density, temperature and other numbers in equation (3) must be taken at an upstream distance $`k_I(E_m)/v_S`$ from the collisionless shock transition. Application to our model yields $`E_m`$ 10 GeV if $`v_{S7}`$ ($`1.11.5`$) and $`k_I(E_m)10^{24}`$ cm<sup>2</sup> s<sup>-1</sup> because of the high ionization in the preshock region (I) and the postshock cooling layer (III) provided by the UV radiation. One can justify the same value of $`E_m`$ for the electrons because the synchrotron and inverse Compton losses are relatively unimportant for $`E_m`$ 10 GeV.
Region I. In the preshock region of the shock with $`v_{S7}>`$ 1.1, we assume a highly ionized plasma of temperature $`T_4`$ 1 up to depths $``$ a few times 10<sup>17</sup> cm<sup>-2</sup> upstream. For the IC 443 preshock region, we adopted a density $`25`$ cm<sup>-3</sup> and $`v_{S7}`$ 1.5 (Fesen & Kirshner 1980). The magnetic field $`B_{}10^5`$ G (Chevalier 1999), while $`B_{}5\times 10^6`$ G (the inclination is assumed to be $`\pi /6`$). The Alfven velocity here is about 5 km s<sup>-1</sup>. We assume a conservative diffusion model in the shock upstream region where $`k_I(p)`$ has $`a`$ = 1 and the moderate scattering โstrengthโ 10 $`\eta _I`$ 100 (Ellison et al. 1999).
Region II. In the shock transition region, injection and heating of the electrons occur due to nonresonant interactions with strong MHD fluctuations generated by the ions. Following the model developed by Bykov & Uvarov (1999) we take
$$D_{II}(p)p^2\overline{C}\left(\frac{\delta B}{B_0}\right)^2\left(\frac{v_a}{v}\right)^2\left(\frac{v}{l_i}\right),$$
(4)
where $`\overline{C}1`$ and $`\delta B>B_0`$. The Alfvenic Mach number of the shock, $`M_a>`$ 10, is moderate for our situation. Nevertheless, the effect of the accelerated particle pressure on the shock structure could be important (see, e.g., Jones & Ellison 1991; Baring et al. 1999). An exact treatment of the effect is not feasible in our model because it would require a kinetic description of the ion injection. We accounted for the effect by correction of the total shock compression ratio $`\delta _t`$ and by introducing an extended modified shock structure with a precursor and a subshock with compression ratio $`\delta _s<\delta _t`$ (e.g., Jones & Ellison 1991). In our case with Mach numbers $`M>`$ 15, the estimated extra compression was about 10% and thus $`\delta _t`$ 4.4. A subshock compression ratio $`\delta _s`$ 3.3 is consistent with the assumption of substantial heating of the precursor gas due to wave dissipation (Berezhko & Ellison 1999). These numbers are in agreement with those from a hybrid simulation of nonlinear shock structure (Giacalone et al. 1997). We found that the resulting high energy emission spectra are sensitive to the particular choice of precursor flow structure in the MeV regime. The keV and GeV emission is not sensitive to the structure if the compression ratios $`\delta _t`$ and $`\delta _s`$ are fixed.
Region III. The postshock cooling layer has about a column density $`3\times `$10<sup>17</sup> cm<sup>-2</sup> of highly ionized plasma with an initial density about $`\delta _t`$ times that of the preshock one. The Alfven velocity in the highly ionized portion of the region (III) is similar to that at the preshock region, $``$ 5 km s<sup>-1</sup>. Large scale vorticity with an amplitude $`<\mathrm{\hspace{0.33em}20}\mathrm{km}\mathrm{s}^1`$ on scales $`\mathrm{\Lambda }<\mathrm{\hspace{0.33em}10}^{14}`$ cm may occur here because of shell instabilities of the radiative shock with $`v_{S7}`$ 1.5 (see, e.g., section 5 in the review by Draine & McKee 1993). This would dominate the low energy electron propagation, providing a spatial diffusion coefficient of $`10^{20}`$ cm<sup>2</sup> s<sup>-1</sup> for the radiative shock parameters described above.
Coulomb losses are important here for electrons with energies below $`E_C`$ 20 keV if the large scale turbulent velocity has a substantial longitudinal component with $`w_{III}20\mathrm{km}\mathrm{s}^1`$. $`E_C>`$ 8 MeV is expected in the absence of a longitudinal component of large scale turbulence in the postshock cooling layer because the Coulomb losses are overcome by resonant interaction with Alfven waves. Since the Alfven velocity is relatively low ($`5\mathrm{km}\mathrm{s}^1`$), the value of $`E_C`$ is much higher in that case. An accurate description of MHD turbulence in the postshock cooling layer is not available now. Thus we considered both limiting cases described above and show the uncertainty introduced by the lack of data concerning the large scale turbulence properties in Fig. 1. To avoid overestimating the nonthermal emission, we used a conservative minimum value of $`E_C`$ = 120 keV for the radiative shock structure described above, although one could expect $`E_C`$ 20 keV in the most favorable case.
Region IV. The density and temperature in the radiative shell depend on the preshock density, magnetic field and abundances (Shull & McKee 1979; Chevalier 1999). From the model of Chevalier (1999) for the above preshock parameters, we adopted $`n`$ 275 cm<sup>-3</sup> ($`N_H1.6\times `$ 10<sup>20</sup> cm<sup>-2</sup>) and $`T`$ 10<sup>2</sup> K in the dense radiative shell. The mass of the dense radiative shell is $`1000`$ $`M_{}`$ and its magnetic field is $`6\times `$ 10<sup>-5</sup> G. The ionization in the dense radiative shell is supported at the level of about a percent by shock accelerated particles penetrating into the shell (see section 3.5 for details). The phase velocity of the high frequency collisionless Alfven waves which are responsible for electron resonant scattering is about 50 km s<sup>-1</sup>; it is determined by the number density of the ions. The Coulomb loss rate of the electrons is then dominated by the neutral component.
We numerically calculated the electron distribution function in regions (I-IV). The results were then used in calculations of the nonthermal emission.
## 3. NONTHERMAL EMISSION FROM IC 443
The SNR IC 443 (G189.1+3.0) is a very good candidate for testing our model because the interaction of the SNR with a molecular cloud was established from radio and infrared line observations (e.g., DeNoyer 1979; Mufson et al. 1986; Burton et al. 1990; van Dishoeck et al. 1993; Richter et al. 1995; Claussen et al. 1997). The high energy fluxes are somewhat better determined than in other cases. Although we concentrate on IC 443, we briefly mention the case of W44, where observations of OH masers and CO clumps again provide convincing evidence of molecular cloud interaction (Claussen et al 1997; Frail & Mitchell 1998).
To model the nonthermal emission from the shell of IC 443 we integrated the local emissivities over the radiative shock structure (regions I $``$IV). As a parcel of gas containing accelerated nonthermal electrons evolves through the radiative shock structure including the dense shell, the local emissivity from bremsstrahlung and inverse Compton was calculated using standard theory (see Appendix A for details). The electron - ion bremsstrahlung emission differential cross sections (Bethe-Heitler formulae) were taken from Akhiezer & Berestetsky (1957) taking into account the Elwert factor, which is important for modeling the keV emission. The electron - electron bremsstrahlung contribution to the $`\gamma `$-ray emissivity was calculated using cross sections derived by Haug (1975). For the inverse Compton emissivity of IC 443, we used the same description of the background photon field as Gaisser et al. (1998). The hard emission production cross sections used in our calculations are similar to those used recently by other authors (Asvarov et al. 1990; Sturner et al. 1997; Gaisser et al. 1998; Baring et al. 1999). We also estimated the contribution of pion decays to the $`\gamma `$-ray emission using a model for the proton component and the cross sections from Dermer (1986). The resulting emission in the EGRET regime was roughly 7 times below that from electron bremsstrahlung, which is consistent with the results of Sturner et al. (1997). We do not discuss the pion component further.
The synchrotron emission of the electrons calculated using the scheme described by Ginzburg (1979) can be compared to the radio observations and to the upper limits at higher frequencies. We also included free-free absorption of low frequency radio waves (dotted line in Fig. 1). The synchrotron losses of a relativistic electron were included following Ginzburg (1979); bremsstrahlung energy losses are relatively unimportant in our case.
### 3.1. Energetic Nonthermal Emission from the Shell
The emission was integrated over the regions of the radiative shock in order to model the spectrum of the entire shell. Region IV is the dominant contributor to the synchrotron and bremsstrahlung emission. The results in Fig. 1 can be compared to the ASCA and EGRET observations of SNRs, as well as to the significant upper limits that have been set at TeV energies by Whipple observations (Buckley et al. 1998) and from the preliminary analysis of CGRO OSSE observations presented by Sturner et al. (1997). We tried different parameter sets to fit the observations.
In Fig. 1, $`\nu F_\nu `$ fits to the observations of IC 443 are presented for the model described in the previous section. The preshock density is taken to be $`n=25\mathrm{cm}^3`$ and the assumed distance to IC 443 is 1.5 kpc (e.g., Fesen & Kirshner 1980). Interstellar photoelectric absorption of the X-ray spectrum is accounted for with a line-of-sight column density $`N_H=2\times 10^{21}`$ cm<sup>-2</sup> using the cross sections from Morrison & McCammon (1983).
For $`v_{S7}`$ = 1.5 in the diffusion model described by equation (2), we used $`k_{III0}=1.1\times 10^{20}`$ cm<sup>2</sup> s<sup>-1</sup>, $`a`$ = 1.0, $`E(p_{}`$) = 1 MeV, and $`E(p_{}`$) = 20 GeV, which are compatible with theoretical estimates of resonant wave-particle interactions. The corresponding value $`\eta `$ 30. While in principle $`\eta `$ can be determined from observations of heliospheric shocks, single spacecraft observations do not provide direct values and are subject to ambiguous interpretation. Values of $`\eta 30`$ are consistent with modeling of the highly oblique solar wind termination shock (Ellison et al. 1999), but smaller values ($`\eta 310`$) have been inferred for interplanetary traveling shocks (Baring et al. 1997). On the other hand, very weak scattering ($`\eta 100`$) has been inferred from observations near corotating interaction regions (e.g., Fisk, Schwadron, & Gloeckler 1997) and in the solar wind when the interplanetary magnetic field is nearly radial (e.g., Mรถbius et al. 1998). We use $`\eta `$ 30 as a conservative value. If strong MHD turbulence is present in the upstream shock region, the $`\eta `$ values could be closer to 1. This might allow an electron acceleration model similar to that described earlier, but for slower shock velocities ($`<\mathrm{\hspace{0.33em}100}\mathrm{km}\mathrm{s}^1`$).
The diffusion propagation model is not unique. We obtained a similar fit to that presented in Fig. 1 for a diffusion model with $`k_{III0}`$ = 10<sup>20</sup> cm<sup>2</sup> s<sup>-1</sup>, $`a`$ = 0.7, $`E(p_{}`$) = 1 MeV, $`E(p_{}`$) = 20 GeV. As mentioned in ยง2, there is an important uncertainty concerning the lack of a quantitative model for the generation of large scale MHD turbulence in the postshock cooling layer. The uncertainty affects the hard X-ray spectral calculations. In view of this, we calculated the expected bremsstrahlung emission for two limiting cases covering the range of uncertainties. Curve 1 in Fig. 1 corresponds to fully developed Alfvenic large scale MHD turbulence in the postshock cooling layer ($`E_C`$ = 120 keV). Curve 2 corresponds to the case of a lack of large scale MHD turbulence ($`E_C`$ = 2 GeV). While the GeV bremsstrahlung emission (as well as synchrotron and inverse Compton emission) is similar in the 2 cases, the hard X-ray emission is sensitive to the uncertainty. We also obtained a similar fit to that presented in Fig. 1 for a model with $`v_{S7}`$ = 1.1 and a diffusion model described by $`k_{III0}=5\times 10^{19}`$ cm<sup>2</sup> s<sup>-1</sup>, $`a`$ = 1.0, $`E(p_{}`$) = 1 MeV, and $`E(p_{}`$) = 10 GeV. The model with $`v_{S7}`$ 1.1 requires a somewhat higher power conversion efficiency than that with $`v_{S7}`$ 1.5 (see ยง3.6), but this value is preferred for the IC 443 radiative shock velocity (Chevalier 1999).
The model of electron injection from the cloud thermal pool described above is based on the ionization structure of a radiative shock, which is very sensitive to the shock velocity if $`v_{S7}`$ 1 (Shull & McKee 1979; Hollenbach & McKee 1989). In this respect, it is instructive to consider a model in which high energy electrons are not related to freshly injected particles, but are reaccelerated cosmic ray electrons (Blandford & Cowie 1982; Chevalier 1999; Cox et al. 1999). To model that case we suppose that the far upstream electron flux is just that observed for Galactic cosmic ray electrons near Earth, but extrapolated back to the energy $`E_{crm}`$ with the same slope as observed above a GeV. The calculations of $`\nu F_\nu `$ for this case are shown in Fig. 2. We consider the radiative shock structure with the parameters described in ยง2 (the same as were used for Fig. 1). The Galactic cosmic ray electron flux in the far upstream region was taken to be 3$`\times 10^2(E/\mathrm{GeV})^3`$ e<sup>ยฑ</sup> s<sup>-1</sup> cm<sup>-2</sup> GeV<sup>-1</sup> sr<sup>-1</sup> (e.g., Berezinsky et al. 1990) for $`EE_{crm}`$. We assumed a flat cosmic ray electron spectrum below $`E_{crm}`$, corresponding to flattening due to Coulomb losses in the interstellar medium. The maximum energy of cosmic ray electrons reaccelerated by the radiative shock is $`E_m`$ 10 GeV (see equation ), as in the case shown in Fig. 1. The energy $`E_{crm}`$ was considered as a free parameter here.
We do not distinguish between electrons and positrons because there is no difference between them in the MHD shock acceleration process considered above and most of the emission from high energy leptons involved in our model remains the same for both kind of particles. Because of interactions with the radiative shock structure, accelerated cosmic ray electrons could provide a good fit for both synchrotron radio and EGRET emission if $`E_{crm}`$ 10 MeV for the radiative shell parameters described above, but fall short for $`E_{crm}`$ 100 MeV.
The relatively low energies, $`E_{crm}`$ 10 MeV, required for the IC443 shell are a potential problem for this model. The radio synchrotron radiation from the Galaxy indicates that the cosmic ray electron spectrum flattens to $`N(E)E^2`$ below energies $`1`$ GeV (Webber 1983). If we use such a spectrum in our model, the nonthermal emission is significantly below that observed. However, there are uncertainties about betatron-type acceleration in the shock compression, about the magnetic field strength, and about the volume of the emitting region. Duin & van der Laan (1975) did successfully model the IC 443 radio flux with a cosmic ray compression model. We found that models with a higher magnetic field and/or larger volume could reproduce the radio flux, but that the $`\gamma `$-ray fluxes observed by EGRET could not be achieved. Cox et al. (1999) have recently concluded that cosmic ray compression by radiative shocks could provide a good fit to radio and gamma-ray emission from W44 if the cosmic ray electron population in the vicinity of the SNR is higher than that of measured in solar neighborhood. They used a somewhat higher magnetic field ($`B`$ 200 $`\mu `$G) in the radiative shell of W44 than the 60 $`\mu `$G field we assumed for the IC 443 shell.
The acceleration time is a constraining parameter in this model because of the relatively low shock speed $`v_{S7}1.5`$ in the dense medium. The test particle shock acceleration time $`t_a(E)`$ can be estimated from the equation
$$t_a=\frac{3}{v_Iv_{II}}_{p_0}^p\left(\frac{k_I(p)}{v_I}+\frac{k_{II}(p)}{v_{II}}\right)\frac{dp}{p}$$
(5)
where the velocities $`v_i`$ are measured in the shock rest frame (e.g., Axford 1981). For the radiative shock parameters described above, the acceleration time for $`E>`$ 10 GeV is about 3,000 years, for $`\eta `$ 30 (note that $`t_a\eta `$), consistent with the age of IC 443 estimated from the hydrodynamical model by Chevalier (1999). On the basis of Ginga observations of IC 443, Wang et al. (1992) suggested an age of about one thousand years and an association with the supernova of AD 837. The reasons for the low age are related to a thermal interpretation of the observed X-ray emission above 10 keV. We shall argue below for the nonthermal origin of the hard emission in a radiative shock model. A distinctive feature of the radiative shock SNR shell model with direct injection of the electrons from the thermal pool is a hard spectrum of nonthermal emission extending from keV to GeV energies (Fig. 1). There is also a hard emission component due to inverse Compton emission of relativistic electrons having a scale comparable to the scale of the whole radio image of the remnant.
### 3.2. Radio Emission from the Shell
Observations of the radio emission from IC 443 were performed with different instruments over the last 35 years (see, e.g., Erickson & Mahoney 1985; Green 1986; Claussen et al. 1997 and references therein). The 151 and 1419 MHz radio maps (Green 1986) reveal that the most intense flux comes from the shell of the remnant. We suggest that the eastern part of the shell is a radiative shock which has a high magnetic field and electron flux, dominating the synchrotron radiation from the remnant.
We calculated the synchrotron radiation from the remnant using the local emissivity given by equation (A17). The integrated spectrum from the remnant is dominated by the emission from the radiative shell (having magnetic field $`B6\times 10^5`$ G) if the magnetic field in the central parts of the remnant (region V) is moderate, $`B4\times 10^6`$ G. For modeling the shell emission, we took into account the interstellar free-free absorption at low radio frequencies, assuming a column density to IC 443 of $`N_H2\times 10^{21}`$ cm<sup>-2</sup> and an average interstellar medium temperature $`T10^3`$ K. This column density is similar to that adopted for the recent analysis of ASCA data by Keohane et al. (1997) and is about four times less than the value used earlier by Petre et al. (1988) and Wang et al. (1992). We assumed that the ionization fraction is $`15`$%. The free-free absorption inside the cool radiative shell is not very important above 50 MHz because the emission measure is about 1 cm<sup>-6</sup> pc. The solid and dot-dashed lines on Fig. 1 show the total synchrotron radiation from IC 443 (dominated by shell emission in our model) expected at the Earth compared to the observational data (Erickson & Mahoney 1985). Both of our models are illustrated: direct injection of thermal electrons (Fig. 1) and reacceleration of preexisting cosmic ray electrons by a radiative shock (Fig. 2). A radiative shock with total compression ratio $`\delta _t`$ = 4.4, including a shock precursor and subshock ($`\delta _s`$ = 3.3), was considered. An important feature of the radio observations of IC 443 and W44 is that their spectral indices are relatively flat, implying electron energy spectra flatter than those produced by standard strong shock acceleration in the test particle limit. In our model, the calculated radio spectrum fits that observed from the IC 443 shell because of the large shock compression ratio $`\delta _t4.4`$ and to a lesser extent because of the effect of second order Fermi acceleration discussed recently by Ostrowski (1999).
### 3.3. Nonthermal Emission from a Shocked Clump
Molecular clouds show complex structure with an ensemble of dense massive clumps embedded in the interclump medium (Blitz 1993). We considered the possibility of radiative shock interaction with dense molecular clumps, resulting in high energy emission. In IC 443 and W44, those regions do not appear to be significant sites of GHz radio continuum emission, but the clumps may contain most of the cloud mass and are possible sites of hard X-rays and MeV $`\gamma `$-ray emission. The interaction of a radiative shell with a molecular clump generates a dense slab bounded by two shocks (Chevalier 1999). The forward shock propagating into the clump is strong (see also the simulations of Jones & Kang 1993). Since the clump size is typically smaller or comparable to that of the radiative shock layer, it is embedded in a pool of nonthermal particles created by the radiative shock according to the scenario described above. Accelerated particles as well as ionizing radiation from the hot interior regions of the SNR should provide ionization of the external portions of a dense clump at the level $`>\mathrm{\hspace{0.33em}1}`$%. This is sufficient for a shock propagating into the dense clump with $`n`$ 10<sup>4</sup> $`\mathrm{cm}^3`$ and $`B`$ 0.1 mG (e.g., Claussen et al. 1997) to be collisionless. The percent ionization level provides Alfven Mach numbers $`M_a>`$ 1 and weak damping of collisionless Alfven waves up to wavelengths $``$ 10<sup>10</sup> cm for a flow with velocity $`>`$$``$ 25 $`\mathrm{km}\mathrm{s}^1`$ in the clump. Then the injection and acceleration processes considered above would be efficient and the ionization in the preshock region would be maintained self-consistently at least at a percent level even for a shock propagating inside the dense clump.
The gas ionization by UV radiation in the radiative shock precursor is sensitive to the shock speed if $`v_{S7}`$ 1.1 (Shull & McKee 1979; Hollenbach & McKee 1989). This implies that a shock having $`v_{S7}`$ 1 propagating in a dense clump would be able to accelerate electrons only below $`E_m<`$ 1 GeV (see equation ). We shall estimate below also the gas ionization rate due to accelerated nonthermal particles. The lifetime of a pre-existing electron of MeV energy against ionization losses in the interclump medium ($`n`$ 10 cm<sup>-3</sup>) is about $`10^4`$ years. Unless a nearby source of fresh accelerated electrons is present, Galactic cosmic ray MeV electrons are unlikely to be present in the cloud. Direct injection of electrons from the thermal pool of the weakly ionized precursor would be a dominant source of shock accelerated electrons up to the cut-off energy $`E_m`$. Because of a lack of accelerated GeV electrons in that case ($`v_{S7}`$ 1) the source would not appear as a bright radio and hard EGRET $`\gamma `$-ray emission feature. Nevertheless, these shocks might manifest themselves as hard X-ray and MeV $`\gamma `$-ray emission sources with $`\nu F_\nu \nu ^a`$ $`(a>\mathrm{\hspace{0.33em}0.5})`$ up to a few MeV and with a soft spectrum at higher energies. A comparison of Fig. 3 with Fig. 1 shows that the hard X-ray flux at 7 keV and above could be dominated by the emission from the shocked clumps, which is consistent with ASCA observations of IC 443 presented by Keohane et al. (1997). The hard X-ray nonthermal emission flux may be variable on a timescale of years. A relatively high MeV emissivity is a distinctive feature of shocked clumps, but they are difficult to observe. They must be nearby (within a few hundreds of parsecs) to be detected with COMPTEL/OSSE CGRO.
For a forward shock of velocity $`v_{S7}`$ 1 propagating into a dense clump of number density $`n`$ 10<sup>3</sup> cm<sup>-3</sup> with upstream ionization level of a few percent (at $`k_I[E_m]/v_S`$ distance), the upper cut-off energy $`E_m<`$ 0.5 GeV. This is about 40 times less than that for a radiative shock in the interclump medium. The spatial diffusion $`k_{i0}`$ below $`p_{}`$ scales as $`w_i\mathrm{\Lambda }`$. Since the scale of a postshock radiative cooling layer $`n^1`$ and $`\mathrm{\Lambda }`$ is a fraction of that, we have $`k_{III0}3\times `$ 10<sup>18</sup> cm<sup>2</sup> s<sup>-1</sup> in the clump case.
The effective compression of the nonthermal electrons as well as the magnetic field by the system of two shocks bounding the slab could be high. The magnetic field in the clump is expected to be high. Claussen et al. (1997) derived line-of-sight magnetic field strengths of $``$0.2 mG, remarkably uniform on the scale of several parsecs. Relativistic electrons accelerated by a shock propagating into a clump should produce radio emission with a flat spectrum up to some hundreds of MHz, which is lower than the maximum radio frequency from the shell because of the lower $`E_m`$ in the clumps. The emission measure of a clump interacting with a radiative shock could be EM $`<`$$``$ 1,000 cm<sup>-6</sup> pc, which is much higher than that of the shell. Thus, the internal free-free absorption could be important here providing a flatter radio spectrum in the 100s of MHz regime. This is possibly a reason for the observed spatial variations of the radio spectral index in IC 443 (e.g., Green 1986).
We present in Fig. 3 the calculated $`\nu F_\nu `$ spectrum of nonthermal emission from a radiative shock - molecular clump interaction region. The forward shock velocity was high โ $`100\mathrm{km}\mathrm{s}^1`$, the initial number density in the clump 1,000 cm<sup>-3</sup>, and the shocked clump radius 0.5 pc. The shock compression ratio was the same as that in the radiative shell model: $`\delta _s`$ 3 and $`\delta _t`$ 4.4. The ionization of gas in the shock upstream was about 5% (at the distances $`k_I[E_m]/v_S`$). To account for some uncertainty in the gas ionization, we present the radio spectra for two possible values of the emission measure EM of the shocked clump: EM = 110 cm<sup>-6</sup> pc (solid line) and EM = 900 cm<sup>-6</sup> pc (dashed line). The diffusion coefficient was $`k_{III0}3\times `$ 10<sup>18</sup> cm<sup>2</sup> s<sup>-1</sup>, $`E(p_{}`$) = 1 MeV. We illustrate the results for two possible maximal energies of the electrons (depending on the time of shock - clump interaction). Curve 1 in Fig. 3 corresponds to $`E_m=`$ 0.5 GeV and $`E_m=`$ 0.05 GeV for curve 2. The synchrotron radio spectrum for curve 2 is below the scale of the plot. The radio spectrum corresponding to curve 1 is very hard because of internal absorption in the shocked clump medium.
We presented above the results for nonthermal emission of a strong radiative shock of $`v_{S7}1`$ propagating into a molecular clump. For the case of radiative shock interaction with the molecular clumps in IC 443, lower shock velocities, $`v_{S7}`$ 0.3, are more realistic (e.g., Cesarsky et al. 1999; Chevalier 1999). The preshock ionization may be dominated mostly by nonthermal particles in such a shock because UV radiation is inefficient. A self-consistent model of MHD collisionless shock propagation into a dense clump requires simultaneous simulations of the MHD turbulence spectral properties and the nonthermal particle kinetics. Instead, we suppose that a collisionless shock transition exists for $`v_{S7}`$ 0.3 and $`n=10^4\mathrm{cm}^3`$ and assume the same diffusion model as described in ยง2. We calculated the upstream gas ionization, finding that the ionized fraction $`x>`$ 0.9 may hold up to depths about 10<sup>15</sup> cm<sup>-2</sup> due to shock accelerated electrons. The Mach number of the shock with $`v_{S7}`$ 0.3 is typically below 5 because of heating of upstream gas up to 10<sup>4</sup> K by accelerated particles and the compression ratio is lower than 4. The maximum energy of accelerated electrons in that case is typically below an MeV. We illustrate the calculated $`\nu F_\nu `$ spectrum of nonthermal emission from a 30 $`\mathrm{km}\mathrm{s}^1`$ velocity shock ($`\delta _t`$ = 3 ) interacting with a molecular clump of density 10$`{}_{}{}^{4}\mathrm{cm}_{}^{3}`$ by curve 3 in Fig. 3; this model corresponds to $`k_{III0}7\times `$ 10<sup>16</sup> cm<sup>2</sup> s<sup>-1</sup> and $`E(p_{}`$) = 10 keV. Such collisionless shocks, if they exist, would provide a softer spectrum of nonthermal radiation than that for high velocity shocks with $`v_{S7}>`$ 1. A more comprehensive study is required to model the nonthermal emission from the low velocity shocks in detail. A search for hard X-ray emission correlated with molecular emission predicted by the simplified model described above is possible with high resolution instruments like Chandra and XMM.
### 3.4. Hard X-ray Emission
IC 443 was a target of X-ray observations with HEAO 1 (Petre et al. 1988), Ginga (Wang et al. 1992), ROSAT (Asaoka & Aschenbach 1994) and ASCA (Keohane et al. 1997). ASCA GIS observations discovered the localized character of the hard X-ray emission (Keohane et al. 1997). Most of the 2-10 keV GIS photons came from an isolated emitting feature and from the southeast elongated ridge of hard emission. The ridge is also coincident with the 95% confidence error circle of the EGRET CGRO source (Esposito et al. 1996), while the isolated feature is outside the circle (Keohane et al. 1997). The integrated 2-10 keV flux observed by ASCA GIS was (5$`\pm 1)\times `$ 10<sup>-11</sup> erg cm<sup>-2</sup> s<sup>-1</sup>, representing more than 90% of the total flux from the SNR (Keohane et al. 1997). The hard X-ray ridge, as defined by Keohane et al. (1997), is only a part of the more extended radio shell discussed in the previous section. The flux density at 7 keV from the isolated emitting feature was about 4$`\times `$ 10<sup>-5</sup> photons cm<sup>-2</sup> s<sup>-1</sup> keV<sup>-1</sup> and that from the ridge was 2$`\times `$ 10<sup>-5</sup> photons cm<sup>-2</sup> s<sup>-1</sup> keV<sup>-1</sup>. HEAO A-2 measured a flux at 2โ10 keV of (7$`\pm 1)\times `$ 10<sup>-11</sup> erg cm<sup>-2</sup> s<sup>-1</sup> (Petre et al. 1988), while Ginga measured a 2โ20 keV flux of about 9$`\times `$ 10<sup>-11</sup> erg cm<sup>-2</sup> s<sup>-1</sup> (Wang et al. 1992). These numbers are consistent with the assumption that the integrated flux is dominated by the extended soft component coming from the central parts of the remnant.
The soft X-ray 0.2โ3.1 keV surface brightness map of IC 443 from the Einstein Observatory (Petre et al. 1988) shows bright features in the northeastern part of the remnant. The presence of nearly uniform X-ray emission from the central part of the remnant is a characteristic feature of mixed-morphology SNRs (Rho & Petre 1998). It might be due to the effect of thermal conduction (Cox et al. 1999; Shelton et al. 1999), although Harrus et al. (1997) argued for an alternative radiative shock model for W44, another mixed-morphology SNR. That feature in our model corresponds to the emission from hot (T$`<\mathrm{\hspace{0.33em}10}^7`$ K), low density gas interior to the shock (region V).
In the radiative shock model with direct injection from the thermal pool and second order Fermi acceleration in postshock cooling layer, most of the hard X-ray emission ($`>\mathrm{\hspace{0.33em}7}`$ keV) comes from the shell - radiative shock structure with a rising $`\nu F_\nu `$ X-ray spectrum (curve 1 in Fig. 1 for the shell and curves 1 and 2 in Fig. 3 for a shocked molecular clump). In addition to the shell, there is a more extended source of hard X-rays from inverse Compton emission. It is the dominant source of hard X-rays in the scenario with cosmic ray electron reacceleration (Fig. 2) and in the case of a lack of second order Fermi acceleration in the radiative shell (see curve 2 in Fig. 1). In the scenario with reacceleration of cosmic ray electrons, it is impossible to obtain the 7 keV flux observed by ASCA from the IC 443 ridge. Thus, a resolved hard X-ray image of the IC 443 remnant may be very informative in determining the electron acceleration scenario.
The isolated emitting feature in the southern part of IC 443 detected by ASCA and brightest around 7 keV, has a flat spectrum low frequency radio continuum source (Green 1986) and is close to a region of excess H<sub>2</sub> emission. High spatial resolution observations of this isolated emitting feature are required to decide between a low-luminosity pulsar nebula and shock interaction models (e.g., radiative shock interaction with a dense molecular clump, Fig. 3). Observations of the hard emission with the sensitivity and resolution available with BeppoSAX, Chandra XRO, XMM and GLAST would be required to constrain the extension and spectral properties of the hard X-ray emission from IC 443.
### 3.5. Gas Ionization by Nonthermal Electrons
Nonthermal particles accelerated by a MHD collisionless shock wave provide an efficient ionizing agent. We calculated above the nonthermal emission generated by accelerated energetic electrons; one can also calculate the ionization produced by the same electrons in the molecular cloud. This might connect the high energy observations with radio, IR and optical emission from shocked atomic and molecular gas. To estimate the gas ionization due to accelerated electrons in the radiative shock structure we used the electron spectra calculated with the kinetic model of electron acceleration and propagation described in ยง2 and electron ionization cross sections from Appendix B. In Fig. 4, we present the calculated ionization rate $`\zeta _e`$ and estimated ionized fraction $`x`$ in the shell of a radiative shock due to accelerated electrons for the case of $`v_{S7}=1.5`$ for both models of large-scale turbulence in the postshock cooling layer considered in ยง3.1.
There should also be contributions to $`x`$ from accelerated nucleons and from UV radiation which are not included in Fig. 4. We accounted for only radiative recombination, assuming a small molecular fraction in the radiative shell at column densities below 2 $`\times 10^{20}`$ cm<sup>-2</sup>.
The solid curves in both panels of Fig. 4 correspond to the case of fully developed Alfvenic turbulence in the postshock cooling layer, which provides efficient second order Fermi acceleration making Coulomb losses relatively unimportant above $`E_C`$ = 120 keV. This case corresponds to curve 1 in Fig. 1 for emission from the shell with substantial hard X-ray emission. The dotted lines in Fig. 4 correspond to the case of a lack of large scale MHD turbulence where the Coulomb losses overcome the second order Fermi acceleration up to $`E_C`$ = 2 GeV. The nonthermal emission expected in this case is illustrated by curve 2 in Fig. 1. The radiative shell ionization by nonthermal electrons is sensitive to the turbulent structure of the postshock cooling layer and there is a correlation between the hard X-ray spectrum and the ionization structure of the radiative shell. We shall consider the effect of nonthermal particles on the ionization structure of a shocked clump elsewhere.
### 3.6. Energy in Nonthermal Electrons
The energy of the electron component for the parameter set used to compute the spectra in Fig. 1 is substantial. The power required for electron acceleration and maintenance for curve 1 in Fig. 1, $`1.5\times 10^{37}`$ erg s<sup>-1</sup>, is high because of the strong Coulomb losses of keV electrons in the dense ionized medium, but it is consistent with the estimate of the total radiative losses from IC 443, dominated by infrared emission (Mufson et al. 1986; Burton et al. 1990). A similar power is required for both models 1 and 2 in Fig. 3, where radiative shock - molecular clump interaction is illustrated. For curve 2, the power requirements are less, $`6\times 10^{36}`$ erg s<sup>-1</sup>. The implied efficiency of power conversion from the MHD shock flow (which can be estimated as 3 $`\times 10^{38}`$ erg s<sup>-1</sup>) to the nonthermal electrons is about 5% in these cases. The model with $`v_{S7}1`$ would provide emission similar to that shown in Fig. 1, but requires about three times higher efficiency. This is higher than the electron acceleration efficiency estimated from GeV regime cosmic rays observed near the Earth, but evolved SNRs interacting with dense molecular gas probably cannot be considered as the main source of the observed Galactic cosmic rays because of the relatively low maximum energies of accelerated particles.
The pressure of the nonthermal electron component downstream of the shock (region IV) is $`๐ซ_e`$ 7$`\times `$ 10<sup>-11</sup> erg cm<sup>-3</sup>. The magnetic pressure in the dense shell is about 1.5 $`\times `$ 10<sup>-10</sup> erg cm<sup>-3</sup>, which is higher than the thermal gas pressure. The uniform ($``$ 60 $`\mu `$G) and stochastic magnetic field components dominate the total pressure in the shell.
Let us consider the energy requirements for the scenario with electron injection from preexisting cosmic rays (Fig. 2). The nonthermal electron pressure in this case is $`๐ซ_e`$ 5.2$`\times `$ 10<sup>-11</sup> erg cm<sup>-3</sup> (for $`E_{crm}`$ 8 MeV), similar to that for the model with injection from the thermal pool, because it is dominated by GeV particles. An advantage of this scenario is that much less power is required. It is about 4 $`\times `$ 10<sup>35</sup> erg s<sup>-1</sup>, which is less than 10% of that for the scenario with injection from the thermal pool. The difference is because Coulomb losses are unimportant for the high energy electrons involved in that scenario. However, a potential problem is the relatively low $`E_{crm}`$ required. The lifetime of a 20 MeV electron against ionization losses in a neutral medium of number density $``$ 1 cm<sup>-3</sup> is about 10<sup>6</sup> years. This implies that the source of MeV cosmic ray electrons should exist within about 100 pc of the molecular cloud if the diffusion coefficient is $`3\times `$ 10<sup>26</sup> cm<sup>2</sup> s<sup>-1</sup> at these energies.
## 4. DISCUSSION AND FUTURE PROSPECTS
For individual supernova remnants, the ambient density is an important parameter that can be estimated from multiwavelength observations (Chevalier 1999). Another important parameter for modeling the nonthermal particles in a SNR in a molecular cloud is the collisionless MHD turbulence spectrum, particularly in the radiative shock cooling layer. A substantial level of MHD collisionless turbulence could overcome Coulomb losses in the dense plasma downstream from the radiative shock. An accurate model of MHD turbulence in the postshock cooling layer is not available now. Thus we considered both limiting cases described in the ยง 2 and show in Fig. 1 the uncertainty introduced by the lack of data concerning the MHD turbulence properties. Coulomb losses are important in the postshock cooling region only for electrons with energies below $`E_C`$ 20 keV if the large scale turbulent velocity has a substantial longitudinal component of $`20\mathrm{km}\mathrm{s}^1`$. A higher value ($`E_C>`$ 8 MeV) is expected in the absence of a longitudinal component of large scale turbulence in the postshock cooling layer because the Coulomb losses are overcome by resonant interaction with Alfven waves. Since the Alfven velocity is relatively low ($`5\mathrm{km}\mathrm{s}^1`$), the value of $`E_C`$ is much higher in that case. We used a conservative minimum value of $`E_C`$ = 120 keV for the radiative shock structure described above, although one could expect $`E_C`$ 20 keV in the most favorable case. The nonthermal emission of SNR in a molecular cloud (especially in the hard X-ray regime) and ionization structure of the radiative shock are sensitive to the MHD turbulence model. Spatially resolved observations of the nonthermal emission from SNRs may be able to constrain models of MHD turbulence.
All of the SNR candidates from the EGRET list: IC 443, $`\gamma `$ Cyg, W44, and Monoceros (Esposito et al. 1996) are old remnants interacting with molecular clouds. The $`\gamma `$-ray emission from electrons accelerated by the radiative shock calculated for the IC 443 parameters presented in Figs. 1 and 2 is in good agreement with that observed by EGRET (Esposito et al. 1996), as well as with the upper limits established by Whipple (Buckley et al. 1998). The high energy $`\gamma `$-rays ($`>\mathrm{\hspace{0.33em}50}`$ MeV) from the radiative shell should be spatially correlated with the radio emission. The EGRET telescope detected an extended excess (95% confidence circle $`42^{}`$ in radius) from the Monoceros SNR correlated with the 1.42 GHz radio emission (Esposito et al. 1996). This is in accordance with the model of a leptonic origin of GeV emission from extended SNRs interacting with clouds (de Jager & Mastichiadis 1997). GLAST (Gamma-ray Large Area Space Telescope) will be an excellent instrument for future $`\gamma `$-ray observations, because it will have the capability of spatially resolving the $`\gamma `$-ray emission from SNRs.
Spatially resolved observations of W44, IC 443, 3C391 and some other mixed-morphology SNRs from the list given by Rho & Petre (1998) with BeppoSAX, Chandra XRO, XMM would be very valuable tools to test our model. The hard X-ray detector (HXD) aboard the forthcoming ASTRO-E mission could be used for observations of the hard 10โ700 keV continuum with a field of view of 0.8 FWHM at 60 keV and 2.8 at 500 keV from the extended hard X-ray emission from IC 443, W44 and the Monoceros SNR predicted by the radiative shock model (see Fig. 1 and Fig. 3). Spatially resolved spectra are needed to distinguish the shell emission from the shocked clump emission.
Molecular clumps interacting with moderately fast radiative shocks are also expected to be sources of hard X-rays and MeV $`\gamma `$-rays (up to 100s of MeV). Nonthermal continuum radio emission (100s of MHz) with a sharply rising $`\nu F_\nu `$ spectrum (Fig. 3) and a time dependent cut-off frequency is expected in the fast shock model. Due to potentially substantial internal free-free absorption, the spectrum of radio emission from a localized clump might constrain the ionized gas density in the clump. The MeV $`\gamma `$-ray spectrum of a localized clump may be resolved with the forthcoming $`\gamma `$-ray missions INTEGRAL (e.g., Schรถnfelder 1999; Winkler 1999) and GLAST. With an expected angular resolution about 12 FWHM, imager IBIS aboard INTEGRAL would allow detection of hard X- ray ( $`>`$$``$ 50 keV) emission from molecular clouds within $``$ 1 kpc. A comparison of hard emission spectra with the radio spectrum can provide valuable information about the density and magnetic field in a clump. A possible variability on a timescale of a few years for clump hard emission and radio emission (on a longer time scale) could further constrain the model.
For low velocity (below $`30\mathrm{km}\mathrm{s}^1`$) shocks interacting with a dense ($`10^4\mathrm{cm}^3`$), magnetized ($`B<`$ 0.5 mG) molecular clump, one may expect hard X-ray emission below the MeV regime correlated with the regions of molecular emission of shocked gas. Nonthermal radio continuum is not expected in the case of low velocity shocks because of a lack of accelerated relativistic electrons at sufficiently high energies.
Finally, we note that the high density of energetic particles in the vicinity of a shock wave in a molecular cloud can affect the ionization and thermal properties of the gas. Rich molecular spectra have been observed from IC 443 and 3C391 (Reach & Rho 1999). It will be interesting to see whether there is a signature of the presence of nonthermal particles that can be discerned from the molecular spectra.
A.M.B thanks Hans Bloemen for useful discussions. We are grateful to M.G. Baring for comments and for making available the corrections to the e-e bremsstrahlung formulae and to the referee for very constructive comments. The work of A.M.B and Yu.A.U was supported by the INTAS grant 96-0390 and that of R.A.C. by NASA grant NAG5-8232.
## Appendix A Appendix A: Radiative Processes
Given the electron distribution function, the high energy emission flux $`J(E_\gamma )`$ at a distance $`R`$ from the source can be calculated from the equation:
$$J(E_\gamma )=\frac{1}{4\pi R^2}\underset{V}{}๐V\frac{dn_\gamma (E_\gamma ,\stackrel{}{r})}{dt}$$
(A1)
where $`dn_\gamma (E_\gamma ,\stackrel{}{r})/dt`$ is the emissivity, which here includes bremsstrahlung, synchrotron radiation and the inverse Compton effect. We generally assumed a pure hydrogen composition in our calculations.
High energy particles penetrating through a partially ionized medium produce photons due to interactions with atoms, electrons and ions. We assumed a Maxwellian distribution for the ambient matter and in most cases took the target ambient particles to be at rest. Then, the emissivity can be calculated from
$$\frac{dn_\gamma (E_\gamma )}{dt}=4\pi \left(\underset{E_\gamma }{\overset{\mathrm{}}{}}๐E_eN_p\sigma _{ep}(E_\gamma ,E_e)J(E_e)+\underset{E_{min}}{\overset{\mathrm{}}{}}๐E_eN_e\sigma _{ee}(E_\gamma ,E_e)J(E_e)\right),$$
(A2)
where $`J(E_e)`$ is the electron flux measured in units s<sup>-1</sup> cm<sup>-2</sup> keV<sup>-1</sup> sr<sup>-1</sup> , $`\sigma _{ep}`$ and $`\sigma _{ee}`$ are differential cross sections of one photon emission due to electron-proton and electron-electron interactions integrated over angles, and $`E_{min}`$ is the minimum energy of incoming electron that could generate a photon with energy $`E_\gamma `$. For e-p interactions, $`E_{min}=E_\gamma `$ is a good approximation. The correction due to finite proton mass is less than the uncertainty in the cross section. In the case of e-e interactions, we used the following equation for $`E_{min}`$:
$$E_\gamma =E_{min}m_ec^2/\left[2m_ec^2+E_{min}\sqrt{E_{min}(E_{min}+2m_ec^2)}\right].$$
The inverse bremsstrahlung contribution is negligible unless there is a very high (far in excess of 100) ratio of protons to electrons.
We used the following cross section approximations in our calculations:
### A.1. e-e bremsstrahlung.
Haug (1975) has obtained a general formula for the e-e bremsstrahlung cross section, but it is more convenient to use simplified approximations. We used relativistic and non-relativistic asymptotic expressions, which were matched in the intermediate energy region (see Baring et al. 1999).
In the relativistic limit, we used approximate formulae given by Baier et al. (1967) with a correction factor from Baring et al. (1999):<sup>1</sup><sup>1</sup>1Equations (A4) and (A6) are somewhat different from that given by Baring et al. (1999). M.G. Baring (private communication) made available the corrections to their original formulae.
$$\sigma _{ee}(E_e,E_\gamma )=(\sigma _1+\sigma _2)A(ฯต_\gamma ,\gamma _e),$$
(A3)
where $`\gamma _e=(E_e+m_ec^2)/m_ec^2`$, $`ฯต_\gamma =E_\gamma /m_ec^2`$,
$$\sigma _1(\gamma _e,ฯต_\gamma )=\frac{4r_o^2\alpha }{ฯต_\gamma }\left[1+\left(\frac{1}{3}\frac{ฯต_\gamma }{\gamma _e1}\right)\left(1\frac{ฯต_\gamma }{\gamma _e1}\right)\right]\left[\mathrm{ln}\left(2(\gamma _e1)\frac{\gamma _e1ฯต_\gamma }{ฯต_\gamma }\right)\frac{1}{2}\right],$$
(A4)
$$\sigma _2=\frac{r_o^2\alpha }{3ฯต_\gamma }\{\begin{array}{cc}[16(1ฯต_\gamma +ฯต_\gamma ^2)\mathrm{ln}\left(\frac{\gamma _e}{ฯต_\gamma }\right)\frac{1}{ฯต_\gamma ^2}+\frac{3}{ฯต_\gamma }4+4ฯต_\gamma 8ฯต_\gamma ^2\hfill & \\ 2(12ฯต_\gamma )\mathrm{ln}(12ฯต_\gamma )(\frac{1}{4ฯต_\gamma ^3}\frac{1}{2ฯต_\gamma ^2}+\frac{3}{ฯต_\gamma }2+4ฯต_\gamma )],\hfill & \hfill ฯต_\gamma \frac{1}{2},\\ \frac{2}{ฯต_\gamma }\left[\left(4\frac{1}{ฯต_\gamma }+\frac{1}{4ฯต_\gamma ^2}\right)\mathrm{ln}(2\gamma _e)2+\frac{2}{ฯต_\gamma }\frac{5}{8ฯต_\gamma ^2}\right],\hfill & \hfill ฯต_\gamma >\frac{1}{2},\end{array}$$
(A5)
$`r_o=e^2/m_ec^2`$ is the classical electron radius, $`\alpha =e^2/\mathrm{}c`$, and
$$A(ฯต_\gamma ,\gamma _e)=1\frac{10}{3}\frac{(\gamma _e1)^{\frac{1}{5}}}{\gamma _e+1}\left(\frac{ฯต_\gamma }{\gamma _e}\right)^{\frac{1}{3}}.$$
(A6)
According to Baring et al. (1999), these formulae agree with those from Haug (1975) with an accuracy $`10\%`$ for electrons of energy $`5`$ MeV .
In the non-relativistic limit, we used equations from Garibyan (1952) and Fedyushin (1952) with correction coefficients from Baring et al. (1999):
$$\sigma _{ee}=\frac{4r_o^2\alpha }{15ฯต_\gamma }F\left(\frac{4ฯต_\gamma }{\gamma _e^21}\right)$$
(A7)
where the photon energy is in the range $`0<ฯต_\gamma <\frac{1}{4}(\gamma _e^21)`$ and $`F(x)`$ ($`0<x<1`$) has the form:
$$\begin{array}{ccc}F(x)\hfill & =\hfill & B(\gamma _e)\left[17\frac{3x^2}{(2x)^2}C(\gamma _e,x)\right]\sqrt{1x}\hfill \\ & & +\left[12(2x)\frac{7x^2}{2x}\frac{3x^4}{(2x)^3}\right]\mathrm{ln}\left(\frac{1+\sqrt{1x}}{\sqrt{x}}\right).\hfill \end{array}$$
(A8)
The correction coefficients are:
$$B(\gamma _e)=1+\frac{1}{2}(\gamma _e^21);C(\gamma _e,x)=\frac{10x\gamma _e\beta _e(2+\gamma _e\beta _e)}{1+x^2(\gamma _e^21)}$$
(A9)
Baring et al. (1999) found less than a $`10\%`$ difference between equation (A7) and the expressions of Haug (1975) for electron energies $`E_e500`$ keV.
### A.2. e-p bremsstrahlung.
In the non-relativistic limit, we took the e-p bremsstrahlung cross section from Akhiezer & Berestetsky (1957):
$$\sigma _{ep}=\frac{16r_0^2Z^2\alpha m_e^2c^2}{3p_1^2}\frac{4\pi ^2\zeta _1\zeta _2}{(e^{2\pi \zeta _1}1)(1e^{2\pi \zeta _2})}\frac{1}{ฯต_\gamma }\mathrm{ln}\left(\frac{p_1+p_2}{p_1p_2}\right),$$
(A10)
which is valid if $`\zeta _11`$; here $`\zeta _i=\alpha Z/\beta _i`$.
For the relativistic limit, we used Born approximation formulae from Akhiezer & Berestetsky (1957):
$$\begin{array}{cc}\sigma _{ep}=\hfill & r_o^2Z^2\alpha \frac{p_2}{p_1}\{\frac{4}{3}2ฯต_1ฯต_2\frac{p_1^2+p_2^2}{p_1^2p_2^2}+m_e(\frac{\eta _1ฯต_2}{p_1^3}+\frac{\eta _2ฯต_1}{p_2^3}\frac{\eta _1\eta _2}{p_1p_2})+\hfill \\ & L[\frac{8}{3}\frac{ฯต_1ฯต_2}{p_1p_2}+\frac{ฯต_\gamma ^2}{p_1^3p_2^3}(ฯต_1^2ฯต_2^2+p_1^2p_2^2)+\frac{m_e^2ฯต_\gamma }{2p_1p_2}(\eta _1\frac{ฯต_1ฯต_2+p_1^2}{p_1^3}\eta _2\frac{ฯต_1ฯต_2+p_2^2}{p_2^3}+\frac{2ฯต_\gamma ฯต_1ฯต_2}{p_1^2p_2^2})]\},\hfill \end{array}$$
(A11)
where $`L=\mathrm{ln}\left(\frac{p_1^2+p_1p_2ฯต_1ฯต_\gamma }{p_1^2p_1p_2ฯต_1ฯต_\gamma }\right)=2\mathrm{ln}\left(\frac{p_1^2+p_1p_2m_e^2}{m_eฯต_\gamma }\right)`$, $`\eta _1=\mathrm{ln}\left(\frac{ฯต_1+p_1}{ฯต_1p_1}\right)=2\mathrm{ln}\left(\frac{ฯต_1+p_1}{m_e}\right)`$, and $`\eta _2=2\mathrm{ln}\left(\frac{ฯต_2+p_2}{m_e}\right)`$. In the velocity range $`Z\alpha (\beta _2^1\beta _1^1)`$, we included the Elwert factor from Koch & Motz (1959): $`f_E=\frac{\beta _1(1exp[2\pi Z\alpha /\beta _1)])}{\beta _2(1exp[2\pi Z\alpha /\beta _2)])}`$.
In equation (A11), $`c=1`$ is assumed. In the intermediate energy range, we matched equation (A10) and equation (A11).
### A.3. Synchrotron radiation.
We used standard formulae for synchrotron radiation in our calculations (e.g., Ginzburg 1979). The emissivity can be written as:
$$\begin{array}{ccc}E_\gamma \frac{dn_\gamma (E_\gamma )}{dt}\hfill & =\hfill & \frac{\sqrt{3}\alpha e}{m_ec}๐\chi ๐EN(E,\chi )H\mathrm{sin}^2(\chi )F\left(\frac{\nu }{\nu _c}\right),\hfill \\ F(x)\hfill & =\hfill & x\underset{x}{\overset{\mathrm{}}{}}K_{5/3}(\eta )๐\eta ,\hfill \end{array}$$
(A12)
where $`\chi `$ is the angle between the electron velocity and the magnetic field, $`N(E,\chi )`$ is the electron distribution function over energy and angle, $`\nu _c=\frac{3eH\mathrm{sin}(\chi )}{4\pi m_ec}\left(\frac{E}{m_ec^2}\right)^2`$, and $`K_{5/3}(x)`$ is the McDonald function. Integrating equation (12) over angle under the assumption of a chaotic magnetic field orientation, we obtain the equation
$$E_\gamma \frac{dn_\gamma (E_\gamma ,\stackrel{}{r})}{dt}=\frac{\sqrt{3}\alpha e}{2\pi m_ec}๐EN(E,\stackrel{}{r})H_{}F\left(\frac{\nu }{\nu _c}\right),$$
(A13)
where $`N(E,\stackrel{}{r})`$ is the electron distribution function as a function of energy and position and $`H_{}=\sqrt{2/3}H(\stackrel{}{r})`$ is the averaged magnetic field.
### A.4. Inverse Compton emission
High energy electrons colliding with photons result in high energy photon production (e.g., Jones 1968; Gaisser et al. 1998; Sturner et al. 1997; Baring et al. 1999). If the electron energy $`E_e\frac{m_e^2c^4}{4E_\gamma }\frac{610^{10}}{E_\gamma (\text{eV})}`$eV ($`E_\gamma `$ is the energy of the emitted photon), then the formulae in the Thompson limit are (e.g., Ginzburg 1979):
$$\frac{dn_\gamma (E_\gamma )}{dt}=\sqrt{3}\pi \sigma _Tm_ec^2๐E_{ph}\sqrt{\frac{E_\gamma }{E_{ph}}}N_{ph}(E_{ph},\stackrel{}{R})J_e(m_ec^2\sqrt{\frac{3E_\gamma }{4E_{ph}}},\stackrel{}{R}),$$
(A14)
where $`\sigma _T=\frac{8\pi }{3}\left(\frac{e^2}{m_ec^2}\right)^2`$ is the Thompson cross section, $`N_{ph}(E_{ph})`$ is the background photon energy spectrum, and $`J_e(E)`$ is the incoming electron flux.
If the electron energy $`E\frac{m_e^2c^4}{4E_\gamma }\frac{610^{10}}{E_\gamma (\text{eV})}`$eV, we must use the Klein-Nishina formulae (integrated over angles):
$$\frac{dn_\gamma (E_\gamma )}{dt}=4\pi ๐E_e๐E_{ph}N_{ph}(E_{ph})\sigma _{KN}(E_\gamma ,E_e,E_{ph})J(E_e),$$
(A15)
$$\sigma _{KN}(E_\gamma ,E_e,E_{ph})=\frac{2\pi r_o^2}{ฯต_{ph}\gamma _e^2}\left[2q\mathrm{ln}(q)+1+q2q^2+\frac{q^2(1q)\mathrm{\Gamma }^2}{2(1+q\mathrm{\Gamma })}\right],$$
(A16)
$$\text{ where }\mathrm{\Gamma }=4ฯต_{ph}\gamma _e,q=\frac{ฯต_\gamma }{(\gamma _eฯต_\gamma )\mathrm{\Gamma }},\text{ }0q1.$$
In the above, $`ฯต_\gamma =E_\gamma /m_ec^2`$, $`ฯต_{ph}=E_{ph}/m_ec^2`$, $`\gamma _e`$ is the electron Lorentz factor, and $`r_o=e^2/m_ec^2`$ is the classical electron radius. The interstellar background photon spectrum, $`N_{ph}(E_{ph},R)`$, was adopted from Mezger et al. (1982) and Mathis et al. (1983) taking into account infrared data from Saken et al. (1992) for the case of IC443 (see also Gaisser et al. 1998).
### A.5. Free-free absorption
Radio waves propagating in the partially ionized thermal plasma are subject to thermal free-free absorption. For $`h\nu kT`$ (the Rayleigh โ Jeans regime), the free-free absorption coefficient $`\alpha _\nu ^{ff}`$ (cm<sup>-1</sup>) can be calculated from
$$\alpha _\nu ^{ff}=\frac{4e^6}{3m_ekc}\left(\frac{2\pi }{3km_e}\right)^{1/2}T^{3/2}Z^2n_en_i\nu ^2\overline{g_{ff}}.$$
(A17)
Using an approximate formula for the Gaunt factor $`\overline{g}_{ff}(\nu ,T)`$ from Rybicki & Lightman (1979) appropriate to our parameter range, we obtained a simplified expression for the optical depth $`\tau (\nu ,T)`$
$$\tau 0.022\nu _{100}^{2.1}T_{100}^{1.34}EM,$$
(A18)
where the radio wave frequency $`\nu _{100}`$ is measured in units of 100 MHz, the plasma temperature in units of 100 K is $`T_{100}`$, and the emission measure $`EM`$ is in units of cm<sup>-6</sup> pc.
## Appendix B Appendix B: Electron Ionization
The primary ionization rate $`\zeta _e`$ due to electron impact was calculated from the equation:
$$\zeta _e=4\pi \underset{E_{min}}{\overset{\mathrm{}}{}}๐E\sigma _{eion}(E)J_e(E),$$
(B1)
where $`J_e(E)`$ is the electron flux measured in units s<sup>-1</sup> cm<sup>-2</sup> keV<sup>-1</sup> sr<sup>-1</sup> , $`\sigma _{eion}`$ is the differential cross section for hydrogen atom primary ionization by electron-atom interactions integrated over atomic electron states, and $`E_{min}`$ is the minimum energy of the ionizing incoming electron.
In the regime below 1 keV, we obtained the following fit to the ionization cross section $`\sigma _{eion}`$ using experimental data from Fite & Brackmann (1958):
$$\sigma _{eion}(E)=\pi a_0^2A_1\mathrm{exp}(bA_4b^3A_5b^5A_6)/E(\mathrm{eV})^{A_3},$$
(B2)
where $`b=A_2/E(\mathrm{eV})`$ and $`E`$(eV) is the incoming electron energy measured in eV. The fitting coefficients are $`A_1`$=76.11, $`A_2`$=14.34, $`A_3`$=0.89, $`A_4`$=3.82, $`A_5`$=โ2.55, and $`A_6`$=4.5. $`a_0=\mathrm{}^2/m_ee^2`$ is the Bohr radius.
At high energies ($`>\mathrm{\hspace{0.33em}1}`$ keV), the Born approximation provides good accuracy (e.g., Mott & Massey 1965):
$$\sigma _{eion}=_0^{\kappa _{max}}_{K_{min}}^{K_{max}}I_{0\kappa }(K)๐K๐\kappa ,$$
(B3)
where
$`I_{0\kappa }(K)`$ $`=`$ $`{\displaystyle \frac{2\pi \mathrm{\hspace{0.17em}2}^{10}\kappa }{\text{}a_0^2k^2K}}{\displaystyle \frac{\mu ^6\left[\text{}K^2+(\mu ^2+\kappa ^2)/3\right]}{\left[\mu ^4+2\mu ^2\left(\text{}K^2+\kappa ^2\right)+\left(\text{}K^2\kappa ^2\right)^2\right]^3}}`$ (B4)
$`\times `$ $`\mathrm{exp}\left[{\displaystyle \frac{2\mu }{\kappa }}\mathrm{arctg}\left({\displaystyle \frac{2\mu \kappa }{K^2\kappa ^2+\mu ^2}}\right)\right]{\displaystyle \frac{1}{1\mathrm{exp}\left(2\pi \mu /\kappa \right)}},`$
$`\mu =Z/a_0`$, $`k(E)`$ is the incoming electron wave number, $`k_\kappa `$ is the outgoing electron wave number, and $`\kappa `$ is the emitted atomic electron wave number. In addition, $`K=|\stackrel{}{k}\stackrel{}{k}_\kappa |`$, $`K_{max}=k+k_\kappa `$, and $`K_{min}=kk_\kappa `$.
We found a convenient approximation providing a reasonably accurate fit (better than 10% above 1 keV and better than 5% above 3 keV) to equations (B3) and (B4):
$$\sigma _{eion}(E)=\pi a_0^2P_1E(eV)^{P_2},$$
(B5)
where $`P_1`$=97.72 and $`P_2`$=0.93.
The gas ionization fraction $`x=n_i/(n_i+n_n)`$ (we consider here a hydrogen plasma) can be expressed through the ionization rate $`\zeta `$ and recombination rate $`\alpha _r(T)`$ as:
$$x=\left(\text{}s^2/4+s\right)^{0.5}s/2$$
(B6)
where $`s=\zeta /(\alpha _rn)`$. We used the radiative recombination rates given by Spitzer (1978) and dissociative recombination rates from Schneider et al. (1994). |
warning/0003/math0003172.html | ar5iv | text | # SQUARE NUMBERS, SPANNING TREES AND INVARIANTS OF ACHIRAL KNOTS This is a preprint. I would be grateful for any comments and corrections!
## 1 Introduction
The main problem of knot theory is to distinguish knots (or links), i.e., smooth embeddings of $`S^1`$ (or several copies of it) into $`^3`$ or $`S^3`$ up to isotopy. A main tool for this is to find invariants of knots, i.e., maps of knot diagrams into some algebraic structure, which are invariant under Reidemeisterโs moves. A family of most popular such invariants are the polynomial invariants, associating to each knot an element in some one- or two-variable (Laurent) polynomial ring over $``$. Given a knot invariant, beside distinguishing knots with it, one is also interested which properties of knots it reflects, and in which way.
One of the most intuitive ways to associate to a knot (or link) another one is to consider its obverse, or mirror image, obtained by reversing the orientation of the ambient space. The knot (or link) is called achiral (or synonymously amphicheiral), if it coincides (up to isotopy) with its mirror image, and chiral otherwise. When considering orientation of the knot, then we distinguish among achiral knots between $`+`$achiral and $``$achiral ones, dependingly on whether the deformation into the mirror image preserves or reverses the orientation of the knot. (For links one has to attach a sign to each component, i.e. embedded circle, and take into account possible permutations of the components.)
When the Jones polynomial $`V`$ \[J\] appeared in 1984, one of its (at that time) spectacular features was that it was (in general) able to distinguish between a knot and its obverse by virtue of having distinct values on both, and (hence) so were its generalizations, the HOMFLY, or skein, polynomial $`P`$ \[F&\] and the Kauffman polynomial $`F`$ \[Ka\]. The $`V`$, $`P`$ and $`F`$ polynomials of achiral knots have the special property to be self-conjugate, that is, invariant when one of the variables is replaced by its inverse. Their decades-old predecessor, the Alexander polynomial $`\mathrm{\Delta }`$ \[Al\], a knot invariant with values in $`[t,t^1]`$, was known always to take the same value on a knot and its mirror image. Nevertheless, contrarily to the common belief, $`\mathrm{\Delta }`$ can also be used to detect chirality (the property of a knot to be distinct from its mirror image) by considering its value $`\mathrm{\Delta }(1)`$, called determinant.
The aim of this paper is to study invariants of achiral knots and to relate some properties of their determinants to the classical topic in number theory of representations of integers as sums of two squares.
In ยง2 we begin with recalling a criterion for the Alexander polynomial of an achiral knot via the determinant, which follows from Murasugiโs work on the signature and the Lickorish-Millett value of the Jones polynomial. This conditions show that, paradoxly formulated, although the Alexander polynomial cannot distinguish between a knot and its mirror image, it can still sometimes show that they are distinct.
After collecting some number theoretic preliminaries in ยง3, we show then in ยง4.1 that the condition of ยง2 is in fact a reduction modulo 36 of the exact arithmetic description of numbers, occurring as determinants of achiral knots. Namely, an odd natural number is the determinant of an achiral knot if and only if it is the sum of two squares. The โonly ifโ part of this statement was an observation of Hartley and Kawauchi in \[HK\]. Our aim will be to show the โifโ part, that is, given a sum of two squares, to realize it as the determinant of an achiral knot (theorem 4.1). The main tool used is the definition of the determinant by means of Kauffmanโs state model for the Jones polynomial \[Ka2\].
Then we attempt to refine our construction, by producing achiral knots (of given determinant) with additional properties: prime and/or alternating. Although it turns out, that one of these properties can always easily be achieved, the situation reveals much harder when demanding them both altogether. We investigate this problem in ยง4.2. Now, the correspondence of ยง4.1 does not hold completely, and there are exceptional values of the determinant, that cannot be realized. To show that 9 and 49 are such, we prove a quadratic improvement of Crowellโs (lower) bound for the determinant of an alternating knot in terms of its crossing number \[C2\], in the case the knot is achiral (proposition 4.2). We obtain then a complete result about which perfect squares can be realized as determinants of prime alternating achiral knots (theorem 4.4).
In ยง4.3 we consider the problem to describe determinants of unknotting number one achiral knots. In this case the description is even less clear, as we show by several examples.
Then we give some applications, including enumeration results on rational knots in ยง5, and a translation of the previously established properties to the number of spanning trees in planar self-dual graphs in ยง6.
Subsequently, in ยง7 we prove some further (at parts still remaining conjectural) properties of the Alexander and skein polynomial of at least large classes of achiral knots, which would allow to decide about chirality (the lack of an isotopy to the mirror image) in a yet different way, at least for these knot classes. These properties concern the leading coefficients of the polynomials, and are closely related to Murasugiโs $``$-product. We prove in particular that perfect squares are exactly the numbers occurring as leading coefficients of the Alexander polynomial of alternating achiral knots, thus improving the previously known necessary condition of non-primeness (corollary 7.1). These results have been obtained with the same arguments independently (but somewhat later) by C. Weber and Q. H. Cรขm Vรขn \[VW\].
Several open problems are suggested during the discussion throughout the paper. These problems appear to be involved enough already for knots, so that we waived on an analogous study of links (which are the cases covering the even natural numbers). For links, also the unpleasant issue of component orientations becomes relevant.
## 2 Detecting chirality with the Alexander polynomial
In the following we will be concerned with the value $`\mathrm{\Delta }(1)`$ of the Alexander polynomial, where $`\mathrm{\Delta }`$ is normalized so that $`\mathrm{\Delta }(t)=\mathrm{\Delta }(1/t)`$ and $`\mathrm{\Delta }(1)=1`$. Up to sign, this numerical invariant can be interpreted as the order of the homology group (over $``$) of the (double) branched covering of $`S^3`$ over $`K`$ associated to the canonical homomorphism $`\pi _1(S^3K)_2`$ and carries the name โdeterminantโ because of its expression (up to sign) as the determinant of a Seifert \[Ro, p. 213\] or Goeritz \[G\] matrix.
To introduce some mathematical notations of the objects thus occurring, let $`D_K`$ be the double branched cover of $`S^3`$ over a knot $`K`$, and let $`H_1(D_K)`$ be its homology group (over $``$). We write then $`det(K)=|\mathrm{\Delta }_K(1)|=|H_1(D_K)|`$.
We start first by description of two special cases of the exact property of the determinant of achiral knots, which we will formulate subsequently, because they have occurred in independent contexts and deserve mention in their own right. They allow to decide about chirality of a knot $`K`$, at least for $`11/18`$ of the possible values of $`\mathrm{\Delta }_K(1)`$.
There is an observation (originally likely, at least implicitly, due to Murasugi \[Mu\], and applied explicitly in \[St\]), using the sign of the value $`\mathrm{\Delta }(1)`$ (with $`\mathrm{\Delta }`$ normalized as said). The information of this sign is equivalent to the residue $`\sigma mod4`$, where $`\sigma `$ denotes the signature. Whenever $`\mathrm{\Delta }(1)<0`$, we have $`\sigma 2mod4`$, so in particular $`\sigma 0`$, and the knot cannot be achiral. This argument works e.g. for the knot $`9_{42}`$ in the tables of \[Ro, appendix\], which became famous by sharing the same $`V`$, $`P`$ and $`F`$ polynomial with its obverse, since its polynomials are all self-conjugate.
Another way to deduce chirality from the determinant is to use the sign of the Lickorish-Millett value $`V\left(e^{\pi i/3}\right)`$ \[LM2\]. Attention to it was drawn in \[Tr\], where it was used to calculate unknotting numbers. Using some of the ideas there, in \[St3\] we observed that this sign implies that if for an achiral knot $`3\mathrm{\Delta }(1)`$, then already $`9\mathrm{\Delta }(1)`$. Thus for example also the chirality of $`7_7`$ can be seen already from its Alexander polynomial, as in this case $`\mathrm{\Delta }(1)=21`$ (although the Murasugi trick does not work here, and indeed $`\sigma =0`$).
Combining both criteria, we arrive in summary to
###### Proposition 2.1
For any achiral knot $`K`$ we have $`\left|\mathrm{\Delta }_K(1)\right|mod36\{1,5,9,13,17,25,29\}`$. $`\mathrm{}`$
An easy verification shows that all these residues indeed occur.
In view of these opportunities to extract chirality information out of $`\mathrm{\Delta }`$, it appears appropriate to introduce a clear distinction between the terms โdetecting chirality of $`K`$โ, meant in the sense โshowing that $`K`$ and $`!K`$ are not the same knotโ (which can be achieved by the above tricks) and โdistinguishing between $`K`$ and $`!K`$โ, meant in the sense โidentifying for a given diagram, known a priori to belong to either $`K`$ or $`!K`$, to which one of both it belongsโ (what they cannot accomplish, but what is the usually imagined situation where some of the other polynomials is not self-conjugate).
Here is a small arithmetic consequence. It is elementary, but is included because of its knot theoretical interpretation and as it is the starting point of exhibiting some more interesting phenomena described in the next sections.
Recall, that a knot $`K`$ is rational (or $`2`$-bridge), if it has an embedding with a Morse function having only four critical points (2 maxima and 2 minima). Such knots were classified by Schubert \[Sh\], and can be alternatively described by their Conway notation \[Co\]. See for example \[Ad, ยง2.3\] for a detailed description. It is well-known that rational knots are alternating (see \[BZ, proposition 12.14, p. 189\]).
###### Corollary 2.1
Let $`p/q`$ for $`(p,q)=1`$, $`p`$ odd be expressible as the continued fraction
$$[[a_1,\mathrm{},a_n,a_n,\mathrm{},a_1]]:=a_1+\frac{1}{a_2+\frac{1}{\mathrm{}a_2+\frac{1}{a_1}}}$$
for a palindromic sequence $`(a_1,\mathrm{},a_n,a_n,\mathrm{},a_1)`$ of even length (with the usual conventions $`\frac{1}{0}=\mathrm{}`$, $`\mathrm{}+=\mathrm{}`$, $`\frac{1}{\mathrm{}}=0`$ for the degenerate cases). Then $`|p|1`$ or $`5mod12`$.
Proof . Observe that $`|p|`$ is the determinant of the achiral rational knot with Conway notation $`(a_1\mathrm{}a_na_n\mathrm{}a_1)`$. The above proposition 2.1 leaves us only with explaining why $`9p`$. The implication $`3det(K)9det(K)`$ for $`K`$ achiral using $`V\left(e^{\pi i/3}\right)`$ follows from the fact that the number of torsion coefficients divisible by $`3`$ of the $``$-module $`H_1(D_K)`$, counted by $`V\left(e^{\pi i/3}\right)`$, is even. However, for a rational knot $`K`$, $`H_1(D_K)`$ is cyclic and non-trivial ($`D_K`$ is a lens space), so that there is only one torsion number at all. Thus $`H_1(D_K)`$ for any achiral rational knot $`K`$ cannot have any $`3`$-torsion. $`\mathrm{}`$
## 3 Number theoretic preliminaries
According to a claim of Fermat, written about 1640 on the margins of his copy of Euclidโs โElementsโ, proved in 1754 by Euler, and further simplified to the length of โone sentenceโ in \[Z3\], any prime of the form $`4x+1`$ can be written as the sum of two squares. More generally, any natural number $`n`$ is the sum of two squares if and only if any prime of the form $`4x+3`$ occurs in the prime decomposition of $`n`$ with an even power, and it is the sum of the squares of two coprime numbers if and only if such primes do not occur at all in the prime decomposition of $`n`$.
The number of representations as the sum of two squares is given by the formula
$$r_2(n):=\frac{1}{4}\mathrm{\#}\{(m_1,m_2)^2:m_1^2+m_2^2=n\}=\underset{dn}{}\left(\frac{4}{d}\right)=\mathrm{\#}\{x:\mathrm{\hspace{0.17em}4}x+1n\}\mathrm{\#}\{x:\mathrm{\hspace{0.17em}4}x+3n\},$$
(1)
which has also an interpretation in the theory of modular forms (see \[HW, (16.9.2) and theorem 278, p. 275\] and \[Z2\]).
A number theoretic explanation of (1) is as follows: If we denote by $`\chi `$ the (primitive) character $`\left(\frac{4}{}\right)`$, with $`4`$ being the discriminant of the field of the Gauร numbers $`[i]`$, we have for $`\mathrm{}(s)>1`$, using that $`[i]`$ has class number $`1`$ and 4 units, that
$$\underset{n=1}{\overset{\mathrm{}}{}}\frac{r_2(n)}{n^s}=\zeta _{[i]}(s)=\zeta (s)L(s,\chi )=\underset{p\text{prime}}{}\frac{1}{(1p^s)(1\chi (p)p^s)},$$
(2)
from which the formula follows by considering the Taylor expansion in $`p^s`$ of the different factors in the product. (This series converges for $`\mathrm{}(s)>1`$.)
The $`\zeta `$-function identities also give a formula for
$$r_2^0(n):=\frac{1}{4}\mathrm{\#}\{(a,b)^2:(a,b)=1,a^2+b^2=n\},$$
(3)
the number of representations of $`n`$ as the sum of squares of coprime numbers.
We have
$$\underset{n=1}{\overset{\mathrm{}}{}}\frac{r_2^0(n)}{n^s}=\frac{\zeta _{[i]}(s)}{\zeta (2s)}=\frac{\zeta (s)L(s,\chi )}{\zeta (2s)}=\underset{p\text{prime}}{}\frac{1p^{2s}}{(1p^s)(1\chi (p)p^s)}=(1+2^s)\underset{\begin{array}{c}p1(4)\\ \text{prime}\end{array}}{}\frac{1+p^s}{1p^s}.$$
(4)
Thus
$$r_2^0(n)=\{\begin{array}{cc}2^k\hfill & \text{if }n=p_1^{k_1}\mathrm{}p_k^{d_k}\text{ or }2p_1^{k_1}\mathrm{}p_k^{d_k}\text{ with }p_1<p_2<\mathrm{}<p_k\text{ primes }1mod4\text{ and }d_i>0\hfill \\ 0\hfill & \text{else}\hfill \end{array}.$$
(5)
Note, that for $`n>1`$,
$$r_2^0(n)=\mathrm{\#}\{(a,b)^2:(a,b)=1,a^2+b^2=n\},$$
(6)
and for $`n>2`$ we have
$$\frac{r_2^0(n)}{2}=\mathrm{\#}\{(a,b)^2:(a,b)=1,ab,a^2+b^2=n\}.$$
(7)
## 4 Sums of two squares and determinants of achiral knots
### 4.1 Realizing sums of two squares as determinants
The aim of this section is to establish a, partially conjectural, correspondence between sums of two squares and the determinant of achiral links. The study of this relation was first initiated in \[HK\], where it was observed that a result of Goeritz \[G\] implies that the determinant of an achiral knot is the sum of two squares. We shall here show the converse. In fact, we have
###### Theorem 4.1
An odd natural number $`n`$ occurs as determinant of an achiral knot $`K`$ if and only if $`n`$ is the sum of two squares $`a^2+b^2`$. More specifically, $`K`$ can be chosen to be alternating or prime, and if one can choose $`a`$ and $`b`$ to be coprime, then we can even take $`K`$ to be rational (or $`2`$-bridge).
Note, that we have given another argument from that of Hartley and Kawauchi for the reverse implication โmodulo 36โ: it was observed above how the signature and the Lickorish-Millett value of the Jones polynomial imply that if $`n`$ is the determinant of an achiral knot, then $`nmod36\{1,5,9,13,17,25,29\}`$. These are exactly the congruences which odd sums of two squares leave modulo 36. Clearly, not every number satisfying these congruences is the sum of two squares. The simplest example is $`77`$. And indeed, this number does not occur as determinant of any achiral knot of $`16`$ crossings.
Additionally to the general case, we also have a complete statement for rational knots. (An analogue for arborescent knots seems possible by applying the classification result of Bonahon and Siebenmann \[BS\].)
###### Theorem 4.2
An odd natural number $`n`$ is the determinant of an achiral rational knot if and only if it is the sum of the squares of two coprime numbers.
The coprimality condition is clearly restrictive โ for example $`49`$ and $`121`$ are not sums of the squares of two coprime numbers. Moreover, it also implies the congruence modulo 12 proved in corollary 2.1.
Fermatโs theorem can be now knot-theoretically reformulated for example as
###### Corollary 4.1
If $`n=4x+1`$ is a prime, then there is a rational achiral knot with determinant $`n`$.
Proof . We have $`n=a^2+b^2`$ and as $`n`$ is prime, $`a`$ and $`b`$ must be coprime. $`\mathrm{}`$
We start by a proof of theorem 4.2. For this recall Krebesโs invariant defined in \[Kr\]. Any tangle $`T`$ can be expressed by its coefficients in the the Kauffman bracket skein module of the room with four in/outputs (see \[St6\]):
$$\text{ }\text{}\text{ }=A\text{ }\text{}\text{ }+B\text{ }\text{}\text{ }.$$
(8)
###### Definition 4.1
For a tangle , we call the numerator closure of $`T`$ and the denominator closure of $`T`$.
When specializing the bracket variable to $`\sqrt{i}`$ ($`i`$ denotes henceforth $`\sqrt{1}`$), $`A`$ and $`B`$ in (8) become scalars.
Then Krebesโs invariant can be defined by
$$Kr(T):=\frac{A}{B}=(A,B)\times /(p,q)(p,q).$$
The denominator and numerator of this โfractionโ give the determinants of the two closures of $`T`$.
Proof of theorem 4.2. A rational achiral knot $`(a_1\mathrm{}a_na_n\mathrm{}a_1)`$ is of the form
$$K=\text{ }\text{}\text{ },$$
(9)
where $`T=(a_1\mathrm{}a_n)`$ is a rational tangle and $`\overline{T}`$ its mirror image. Because of connectivity reasons $`T`$ must be of homotopy type or , i.e. $`Kr(T)=[[a_1,\mathrm{},a_n]]=\frac{p}{q}`$ with $`(p,q)=1`$ and exactly one of $`p`$ and $`q`$ is odd. Thus $`K`$ is the numerator closure of the tangle sum $`T+\overline{T}^T`$, where $`.^T`$ denotes transposition. By the calculus introduced by Krebes, his invariant is additive under tangle sum, and invertive under transposition, and so we have
$$\frac{det(K)}{}=Kr(T+\overline{T}^T)=Kr(T)+\frac{1}{Kr(T)}=\frac{p}{q}+\frac{q}{p}=\frac{p^2+q^2}{pq}.$$
To justify our choice of sign in this calculation, that is, that the determinant is $`p^2+q^2`$ rather than $`p^2q^2`$, it suffices to keep in mind that the diagram (9) is alternating and in calculating the bracket of alternating diagrams no cancellations occur, as explained also in \[Kr\]. Thus we have the โonly ifโ part.
For the โifโ part note that if $`a`$ and $`b`$ are coprime, then $`\frac{a}{b}`$ can be expressed by an continued fraction, and hence as $`Kr(T)`$ for some rational tangle $`T`$. Then $`a^2+b^2`$ (with the above remark on signs) is the determinant of the achiral knot shown in (9). $`\mathrm{}`$
Now we modify the second part of the proof to deduce theorem 4.1. In the following we use Conwayโs notation for tangle sum and product. (See for example again \[Ad, ยง2.3\] for a detailed description.)
Proof of theorem 4.1. Let $`n=p^2+q^2`$. If $`q=0`$ then $`K=T(2,p)\mathrm{\#}T(2,p)`$ ($`T(2,p)`$ denoting the $`(2,p)`$-torus knot) is an easy example, so let $`q0`$. Krebes shows that for any pair $`(p,q)`$ with at least one of $`p`$ and $`q`$ odd there is a(n arborescent) tangle $`T`$ with $`Kr(T)=\frac{p}{q}`$. In fact, $`T`$ can be chosen to be the connected sum of a rational tangle and a knot of the type $`T(2,p)`$ (which can be done in a way the tangle remains alternating). Then again consider the knot in (9) (it is a knot because of the parities of $`p`$ and $`q`$), and from the proof of theorem 4.2 one sees that it has the desired determinant $`n`$.
The knots constructed in (9) then are all alternating. It remains to show that they can be made prime (possibly sacrificing alternation). If $`(p,q)=1`$, then $`K`$ is rational, and hence prime. Thus let $`n=(p,q)>1`$. Then we can choose $`T`$ to be
$$T=\text{ }\text{}\text{ },$$
(10)
that is, in Conwayโs notation $`T=T^{}(0,n)`$ with a tangle $`T^{}`$ being a rational tangle $`a^{}/b^{}`$ with $`a^{}=p/n`$, $`b^{}=q/n`$.
Now replace the $`0`$-tangle in (10) by the (flipped) โKT-grabberโ tangle $`KT`$ in \[Bl, Fl\]. By the same argument as in \[Fl\] (or see also \[KL, Va\]), the tangle $`T_1=T^{}(KT0,n)`$ becomes prime, and hence so is then the knot $`K_1=\overline{T_1T_1}`$ by proposition 1.3 of \[Bl\] (bar denotes tangle closure). As in \[Bl\], $`K_1`$ and $`K`$ have the same Alexander polynomial, so in particular the same determinant. $`\mathrm{}`$
###### Example 4.1
To demonstrate the elegance of theorem 4.1 as a chirality criterion, we remark that among the prime knots of $`10`$ crossings (denoted henceforth according to Rolfsenโs tables \[Ro, appendix\]) there are 6 chiral knots with self-conjugate HOMFLY polynomial โ $`9_{42}`$, $`10_{48}`$, $`10_{71}`$, $`10_{91}`$, $`10_{104}`$ and $`10_{125}`$, and this method shows chirality of four of them โ $`9_{42}`$, $`10_{71}`$, $`10_{104}`$ and $`10_{125}`$, including the two examples ($`9_{42}`$ and $`10_{71}`$) where additionally even the Kauffman polynomial is self-conjugate. (For $`9_{42}`$ and $`10_{125}`$ the congruence modulo 4 is violated, so that, as remarked on several other places, the signature works as well.)
###### Remark 4.1
Since slice knots have square determinant, it also follows that if there exists a rational knot $`S(p,q)`$ which is at the same time achiral and slice, then it will correspond to a Pythagorean triple, that is, be of the Schubert form $`S((m^2+n^2)^2,2mn(m^2n^2))`$ with $`m`$ and $`n`$ coprime.
With regard to theorem 4.1, we conjecture an analogous statement to hold for links.
###### Conjecture 4.1
An even natural number $`n`$ occurs as determinant of an achiral link if and only if $`n`$ is the sum of two squares.
As a remark on links, note that by the above description of numbers which are sums of two squares, this set is closed under multiplication, corresponding on the level of determinants of links to taking connected sums. Thus it would suffice to prove conjecture 4.1 just for prime links.
More number theoretic results on the square representations (which by the said above can also be transcribed knot-theoretically) may be found in \[K, W\].
### 4.2 Determinants of prime alternating achiral knots
Theorem 4.1 naturally suggests the question in how far the properties alternation and primeness can be combined when realizing a sum of two squares as determinant of achiral knots.
In this case, the situation is much more difficult, though. It is easy to see that not every (odd) sum of 2 squares can be realized. The first (and trivial) example ist $`1`$, since the only alternating knot with such determinant is the unknot \[C2\], and it is by definition not prime. However, there are further examples.
###### Proposition 4.1
Let $`n=9`$ or $`n=49`$. Then there is no prime alternating achiral knot of determinant $`n`$.
By the above cited result of Crowell, one has a bound on the crossing number of an alternating knot of given determinant, so could check for any $`n`$ in finite time whether it is realized or not. This renders the check for $`n=9`$ easy. However, the estimate we obtain from Crowellโs inequality is intractable in any practical sense for $`n=49`$. We give an improvement of Crowellโs result for achiral alternating links, which, although not completely sharp, is enough for our purpose.
For the understanding and the proof of this result we recall some standard terminology for knot diagrams.
###### Definition 4.2
The diagram on the left of figure 1 is called connected sum $`A\mathrm{\#}B`$ of the diagrams $`A`$ and $`B`$. If a diagram $`D`$ can be represented as the connected sum of diagrams $`A`$ and $`B`$, such that both $`A`$ and $`B`$ have at least one crossing, then $`D`$ is called disconnected (or composite), else it is called connected (or prime). Equivalently, a diagram is prime if any closed curve intersecting it in exactly two points, does not contain a crossing in one of its in- or exterior.
If a diagram $`D`$ can be written as $`D_1\mathrm{\#}D_2\mathrm{\#}\mathrm{}\mathrm{\#}D_n`$, and all $`D_i`$ are prime, then they are called the prime (or connected) components/factors of $`D`$.
###### Definition 4.3
The diagram is split, if there is a closed curve not intersecting it, but which contains parts of the diagram in both its in- and exterior.
By \[Me\] an alternating link is prime/split iff any alternating diagram of it is so.
###### Definition 4.4
A crossing $`q`$ in a link diagram $`D`$ is called nugatory, if there is a closed (smooth) plane curve $`\gamma `$ intersecting $`D`$ transversely in $`q`$ and nowhere else. A diagram is called reduced if it has no nugatory crossings.
By \[Ka2, Mu3, Th\], each alternating reduced diagram is of minimal crossing number (for the link it represents).
###### Definition 4.5
A flype is a move on a diagram shown in figure 4.
By the fundamental work of \[MT\], for two alternating diagrams of the same alternating link, there is a sequence of flypes (and $`S^2`$-moves) taking the one diagram into the other.
###### Definition 4.6
To define the sign of a crossing in a link diagram, choose an orientation of the link. The sign is then given as follows:
$$\begin{array}{cc}\text{ }\text{}\text{ }& \text{ }\text{}\text{ }\\ +& \end{array}.$$
(11)
A crossing of sign $`+`$ we call positive, and a crossing of sign $``$ we call negative. Note, that the definition requires a link orientation, but for a knot it is independent on which of its both possible choices is taken.
The writhe is the sum of the signs of all crossings in a diagram. It is invariant under simultaneous reversal of orientation of all components of the diagram, so is in particular well-defined for unoriented knot diagrams. It may, however, change if some (but not all) components of a link diagram are reverted.
Crowellโs result about the determinant of alternating links is
###### Theorem 4.3
(Crowell) If $`L`$ is a non-split alternating link of $`n`$ crossings, then $`det(L)n`$, and if $`L`$ is not the $`(2,n)`$-torus link, then $`det(L)2n3`$.
We will show
###### Proposition 4.2
If $`L`$ is an alternating non-split achiral link of $`2n`$ crossings, then $`det(L)n(n3)`$.
Since we use the checkerboard colorings for the proof, our result holds for the most general notion of achirality for links โ we allow the isotopy taking a link to its mirror image to interchange components and/or preserve or reverse their orientations in an arbitrary way. We will define, however, checkerboard colorings later, in ยง6, so that we defer the proof of proposition 4.2 to that later stage.
###### Remark 4.2
The condition the crossing number of $`L`$ to be even is no restriction. The crossing number of any alternating achiral (in the most general sense, as remarked after Proposition 4.2) link diagram is even by \[MT\], since flypes preserve the writhe of the alternating diagram, and reversal of any single component alters the writhe by a multiple of $`4`$ (any two components have an even number of common crossings by the Jordan curve theorem). Thus the writhe must be even, and hence so must be the crossing number.
Proof of proposition 4.1. This is now feasible. Check all the alternating achiral knots in the tables of \[HT\] up to 16 crossings. $`\mathrm{}`$
There is some possibility that the values of proposition 4.1 are indeed the only exceptions.
###### Conjecture 4.2
Let $`n`$ be an odd natural number. Then $`n`$ is the determinant of a prime alternating achiral knot if and only if $`n`$ is the sum of two squares and $`n\{1,9,49\}`$.
At least this is true up to $`n2000`$. A full confirmation of this conjecture is so far not possible, but we obtain a complete statement for $`n`$ being a square.
###### Theorem 4.4
Let $`n`$ be an odd square. Then $`n`$ is the determinant of a prime alternating achiral knot if and only if $`n\{1,9,49\}`$.
Recall, that a knot $`K`$ is called strongly achiral, if it admits an embedding into $`S^3`$ pointwise fixed by the (orientation-reversing) involution $`(x,y,z)(x,y,z)`$. Again dependingly on the effect of this involution on the orientation of the knot we distinguish between strongly $`+`$achiral and strongly $``$achiral knots.
Let $`=[t,t^1]`$ be the Laurent polynomial ring in one variable. For $`F,G`$ write $`FG`$ if $`F`$ and $`G`$ differ by a multiplicative unit in $`[t,t^1]`$, that is, $`F(t)=\pm t^nG(t)`$ for some $`n`$.
The result of \[HK\] is the following.
###### Theorem 4.5
(\[HK\]) If $`K`$ is strongly negative amphicheiral, then $`\mathrm{\Delta }(t^2)F(t)F(t)`$ for some $`F`$ with $`F(t)F(t^1)`$. If $`K`$ is strongly positive amphicheiral, then $`\mathrm{\Delta }(t)=F(t)^2`$ for some $`F`$ with $`F(t)F(t^1)`$.
This theorem will not be used here, but provides some heuristics for the proof of theorem 4.4, and will come to more detailed mention later, so it is possibly appropriate to introduce it here.
Proof of theorem 4.4. Since $`n`$ is a square, it is suggestive by the result of \[HK\] to consider strongly $`+`$achiral knots as candidates to realize $`n`$ as determinant. We consider diagrams of the type
$$D(T_1)=\text{ }\text{}\text{ }$$
Define a pairing $`<T_1,T_2>`$ on the diagram algebra $`DS_3(A=\sqrt{i})`$ (see \[Ka\]) by table 1 (compare also to the pairing $`<,>_3`$ in ยง4 of \[St5\]).
Then $`det(D(T_1))=<T_1,T_1>`$. Let $`T`$ be a tangle
with $`T_1=A\text{ }\text{}\text{ }+B\text{ }\text{}\text{ }`$ and $`T_2=X\text{ }\text{}\text{ }+Y\text{ }\text{}\text{ }`$. Then
$$T=a\text{ }\text{}\text{ }+b\text{ }\text{}\text{ }+c\text{ }\text{}\text{ }+d\text{ }\text{}\text{ }+e\text{ }\text{}\text{ }$$
with
$`a`$ $`=`$ $`XA`$
$`b=d`$ $`=`$ $`BX`$
$`c`$ $`=`$ $`AX+BY+AY`$
$`e`$ $`=`$ $`BY,`$
and
$$<T,T>=\left[(X+Y)(A+B)\right]^2.$$
Whenever $`(X,Y)`$ and $`(A,B)`$ are relatively prime, and $`X,Y,A,B>0`$, one can substitute rational tangles for $`T_{1,2}`$ obtaining a prime alternating diagram of a strongly $`+`$achiral knot. Setting $`X=B=1`$ and varying $`Y`$ and $`A`$, we see that we can cover all cases when $`n=p^2`$ with $`p`$ composite.
Since we dealt with $`p=1`$, it remains to consider $`p`$ prime. If $`p1mod4`$, then (1) shows that $`n`$ has a non-trivial representation as sum of two squares, which then must be coprime. In this case there is an achiral rational knot of determinant $`n`$.
Thus assume $`n=p^2`$ with $`p3mod4`$ prime. We show now that almost all (not necessarily prime) $`p3mod4`$ can be realized.
Consider diagrams $`D(T)`$ for $`T`$ of the form
with $`T_1=X\text{ }\text{}\text{ }+Y\text{ }\text{}\text{ }`$, $`T_2=A\text{ }\text{}\text{ }+B\text{ }\text{}\text{ }`$ and $`T_3=C\text{ }\text{}\text{ }+D\text{ }\text{}\text{ }`$. We find after multiplying out the polynomial and some manipulation
$$<T,T>=\left[X(DA+BC)+Y(BD+AC)\right]^2.$$
Set $`X=1`$ and let $`Y=k`$ vary. The rest is done by choosing small special values for $`A,B,C,D`$.
$$\begin{array}{cccccccccc}& D& B& A& C& DA+BC& BD+AC& \sqrt{<T,T>}& T_2& T_3\\ & & & & & & & & & \\ \text{}\text{}& 2& 3& 2& 1& 7& 8& 7+8k& \text{ }\text{}\text{ }& \text{ }\text{}\text{ }\\ \text{}& 2& 7& 2& 1& 11& 16& 11+16k& \text{ }\text{}\text{ }& \text{ }\text{}\text{ }\\ \text{}& 4& 3& 4& 1& 19& 16& 19+16k& \text{ }\text{}\text{ }& \text{ }\text{}\text{ }\end{array}$$
Examples of the knots thus obtained are given in figure 3 (for simplicity, just the plane curves are drawn).
All diagrams (and hence knots \[Me\]) are prime for $`k1`$. Thus the only cases remaining to check are for $`\sqrt{n}=p\{3,7,11,19\}`$. For $`p=11`$ we have $`10_{123}`$, and for $`p=19`$ we check the knots in the tables of \[HT\]. We obtain the examples $`12_{1019}`$ (the closure of the $`5`$-braid $`(\sigma _1\sigma _2^1\sigma _3\sigma _4^1)^3`$, with $`\sigma _i`$ being the Artin generators, as usual) and $`14_{18362}`$ (the closure of the $`3`$-braid $`\sigma _1^2\sigma _2^3\sigma _1^2\sigma _2^2\sigma _1^3\sigma _2^2`$). The cases $`p=3,7`$ were dealt with in proposition 4.1. $`\mathrm{}`$
###### Remark 4.3
We showed that in fact we can realize any $`n`$ stated in the theorem by a strongly $`+`$achiral or rational knot, so in particular by a strongly achiral knot, since an achiral rational knot is known to be strongly $``$achiral (see \[HK\]). It may be possible to exclude rational knots when allowing the further exception $`n=25`$.
To examine the general case of $`n`$ (not only perfect squares), one needs to consider larger series of examples. Because of \[HK\] the knots should not (only) be strongly $`+`$achiral. A natural way to modify the examples in the proof of theorem 4.4 to be $``$achiral is to consider (braid type) closures of tangles like
$$\text{ }\text{}\text{ }.$$
We just briefly discuss this series to explain some of the occurring difficulties.
Using the Kauffman bracket skein module coefficients
$`=`$ $`A\text{ }\text{}\text{ }+B\text{ }\text{}\text{ },`$
$`=`$ $`C\text{ }\text{}\text{ }+D\text{ }\text{}\text{ },`$
$`=`$ $`X\text{ }\text{}\text{ }+Y\text{ }\text{}\text{ },`$
one finds an expression for $`det(K)`$ as polynomial in $`A,B,C,D,X,Y`$ as before, and after some manipulation arrives at $`det(K)=f_1^2+f_2^2`$ with
$$f_1(A,B,C,D,X,Y)=X(AD+BC)+Y(AC+BD)\text{and}f_2(A,B,C,D,X,Y)=Y(ADBC).$$
(12)
(The correctness of the square decomposition is straightforward to check, but for finding it it is helpful to notice that the substitutions $`Y=0`$ and $`A=C,B=D`$ turn the knots into strongly $`+`$achiral ones, which have square determinant.)
One can conclude from (12) that no number of the form $`n=5p^2`$ with $`p3mod4`$ prime can be written as $`f_1^2+f_2^2`$ with $`f_{1,2}`$ as in (12) for $`A,B,C,D,X,Y`$, unless $`(A,B)`$, $`(C,D)`$ or $`(X,Y)`$ is one of $`(0,p)`$, $`(p,0)`$ or $`(p,p)`$. However, no alternating arborescent tangle has such pair of Kauffman bracket skein module coefficients. Therefore, the above series cannot realize these determinants. On the other hand, the small cases in it (for $`p11`$) are realized by knots with Conway polyhedron $`8^{}`$.
Then one can consider more patters and write down more complicated polynomials, each time having to show that each (at least sufficiently large) number is realized by (at least some of) these polynomials. Presently, such problems in number theory seem very difficult. (One classic example is the determination of the numbers $`G(n)`$ and $`g(n)`$ in Waringโs problem, see for example \[DHL, Ho, HW\].) Therefore, conjecture 4.2 may be hard to approach as of now.
### 4.3 Determinants of unknotting number one achiral knots
We conclude our results on sums of two squares by a related, although somewhat auxiliary, consequence of the unknotting number theorem of Lickorish \[Li\] and its refined version given in \[St3\].
Let $`u_\pm `$ denote the signed unknotting number, the minimal number of switches of crossings of a given sign to a crossing of the reversed sign needed to unknot a knot, or infinity if such an unknotting procedure is not available (this is somewhat different from the definition of \[CL\]).
Thus a knot $`K`$ has $`u_+(K)=1`$ (resp. $`u_{}(K)=1`$) if it can be unknotted by switching a positive (resp. negative) crossing in some of its diagrams.
###### Proposition 4.3
Let $`K`$ be a knot with $`u_+=u_{}=1`$ (for example, an achiral unknotting number one knot). Then $`det(K)`$ is the sum of two squares of coprime numbers.
Proof . Clearly $`\sigma (K)=0`$, so that any of the relevant crossing changes does not alter the signature, and then by the refinement of Lickorishโs theorem given in \[St3\], we have $`\lambda (g_\pm ,g_\pm )=\pm 2/det(K)/`$ for some generators $`g_\pm `$ of $`H_1(D_K)`$. Here $`\lambda `$ is the linking form on $`H_1(D_K)`$ (whose order, as discussed in the introduction, is given by $`|\mathrm{\Delta }_K(1)|`$ ). Thus $`2l^2=2h^2`$ for some $`l,h_{det(K)}^{}`$. Then this group possesses square roots of $`1`$.
The structure of the group $`_n^{}`$ of units in $`_n=/n`$ is known; see e.g. \[Z, exercise 1, ยง5, p. 41\]. From this structure and (5), we see that the number of square roots of $`1`$ in $`_n^{}`$ (for $`n>1`$ odd) is identical to $`r_2^0(n)`$. (D. Zagier remarked to me that one can in fact give a natural bijection between the square roots of $`1`$ in $`_n^{}`$ and representations of $`n`$ as sum of coprime squares.) $`\mathrm{}`$
There are also several questions opened by proposition 4.3. Having the inclusions
$$\{det(K):K\text{ achiral, }u(K)=1\}\{det(K):u_+(K)=u_{}(K)=1\}\{a^2+b^2:(a,b)=1,2a+b\}$$
(to introduce some notation, let $`S_a`$, $`S_\pm `$ and $`S`$ denote these three sets from left to right), the first question is whether and/or which one of these inclusions is proper (or not).
This seems much more difficult to decide than the proof of theorems 4.1 and 4.2. There is no such straightforward procedure available to exhaust all values in $`S`$, and to show a proper inclusion one will face the major problem of deciding about unknotting number one.
After a computer experiment with the knot tables and tools available to me, the smallest $`xS`$ I could not decide to belong to $`S_\pm `$ is 349, whereas the smallest possible $`xS`$ with $`xS_a`$ is only 17. Contrarily to theorems 4.1 and 4.2, very many entries have been completed only by non-alternating knots (which have smaller determinant than the alternating ones of the same crossing number), and in fact we can use the number 17 to show that for this problem non-alternating knots definitely need to be considered.
###### Example 4.2
We already quoted Crowellโs result, that for a given crossing number $`n`$, the $`n`$ crossing twist knot has the smallest determinant $`2n3`$ among alternating knots $`K`$ of crossing number $`n`$, except if $`n`$ is odd, in which case the only knot of smaller determinant is the $`(2,n)`$-torus knot. (A more modern proof can be given for example by the Kauffman bracket, similarly to proposition 4.2.) Thus, except for the $`(2,17)`$-torus knot, any alternating knot of determinant 17 has $`10`$ crossings. A direct check shows that the only such knots of $`\sigma =0`$ are $`8_3`$ and $`10_1`$. However, $`u(8_3)=2`$ as shown by Kanenobu-Murakami \[KM\], and that $`10_1`$ cannot simultaneously have $`u_+=u_{}=1`$ follows by refining their method (see \[St7\]). Thus there is no alternating knot of determinant 17 with $`u_+=u_{}=1`$, and the inclusion
$$S_{a\pm }:=\{det(K):K\text{ alternating, }u_+(K)=u_{}(K)=1\}S$$
is proper. (Contrarily, there is a simple non-alternating knot, $`9_{44}`$, with $`u_+=u_{}=1`$ and determinant 17.) Is it infinitely proper, i.e., is $`|SS_{a\pm }|=\mathrm{}`$?
It is suggestive that considering the even more restricted class of rational knots, the inclusions are infinitely proper. We confirm this for achiral unknotting number one rational knots.
###### Proposition 4.4
$`\left|S\{det(K):K\text{ achiral and rational, }u(K)=1\}\right|=\mathrm{}`$. (In fact, this set contains infinitely many primes.)
Proof . These knots were classified in \[St4, corollary 2.2\] to be those with Conway notation $`(n11n)`$ and $`(3(12)^n1^4(21)^n3)`$. It is easy to see that therefore the inclusion
$$\{det(K):K\text{ achiral and rational, }u(K)=1\}S$$
is infinitely proper. For example, the determinant of both series grows quadratically resp. exponentially in $`n`$, so that
$$\underset{\begin{array}{c}K=S(p,q)\\ \text{achiral, }u(K)=1\end{array}}{}\frac{1}{p}<\mathrm{},$$
while $`{\displaystyle \underset{pS}{}}{\displaystyle \frac{1}{p}}=\mathrm{}`$. (By Dirichlet already $`{\displaystyle \underset{\begin{array}{c}p1(4)\\ \text{prime}\end{array}}{}}{\displaystyle \frac{1}{p}}=\mathrm{}`$ , see \[Z, Korollar, p. 46\].) $`\mathrm{}`$
It appears straightforward to push the method of \[St4\] further to show the same also for rational knots with $`u_+=u_{}=1`$ (although I have not carried out a proof in detail). By refining Kanenobu-Murakami, we first show that there is a crossing of the same sign unknotting the alternating diagram of the rational knot (see \[St7\]). Then consider alternating rational knot diagrams with two unknotting crossings (it does not even seem necessary to have them any more of different sign), and apply the same argument as above, using \[St4, corollary 2.3\] and the remark after its proof.
## 5 Enumeration of rational knots by determinant
The results of ยง2 can be used to enumerate rational knots by determinant. We have for example:
###### Proposition 5.1
The number $`c_n`$ of achiral rational knots of given determinant $`n`$ is given by
$$c_n=\{\begin{array}{cc}1/2r_2^0(n)\hfill & \text{ if }n>2\text{ odd,}\hfill \\ 0\hfill & \text{ else\hspace{0.17em}.}\hfill \end{array}$$
Proof . Use the fact that there is a bijective correspondence between the rational tangle $`T`$ in a diagram (9) of an achiral rational knot $`K`$ and its Krebes invariant $`p/q`$ (with $`pq`$ and $`(p,q)=1`$) giving $`det(K)=p^2+q^2`$. $`\mathrm{}`$
From (5) we obtain then
###### Corollary 5.1
The number of achiral rational knots of given determinant $`n`$ is either zero or a power of two. $`\mathrm{}`$
As a practical application of the argument in the argument proving proposition 5.1 we can consider the achiral rational knots $`(1\mathrm{}1)`$ and $`(31\mathrm{}13)`$ (with the number list of even length) and the tangles $`T`$ obtained from the halves of the palindromic sequence. This way one arrives to a knot theoretical explanation of the identities
$$F_{2n+1}=F_n^2+F_{n+1}^2\text{and}L_{2n+1}+2L_{2n}=L_n^2+L_{n+1}^2,$$
(13)
where $`F_n`$ is the $`n`$-th Fibonacci number ($`F_1=1`$, $`F_2=1`$, $`F_n=F_{n1}+F_{n2}`$) and $`L_n`$ is the $`n`$-th Lucas number ($`L_1=2`$, $`L_2=1`$, $`L_n=L_{n1}+L_{n2}`$). Thus we have
###### Proposition 5.2
There are achiral rational knots with determinant $`F_{2n+1}`$ and $`L_{2n+1}+2L_{2n}`$ for any $`n`$, or equivalently, any (prime) number $`4x+3`$ does not divide $`F_{2n+1}`$ and $`L_{2n+1}+2L_{2n}`$. $`\mathrm{}`$
For $`F_{2n+1}`$ this is a task I remember from an old issue of the Bulgarian journal โMatematikaโ. Recently I found (by electronic search) that it was conjectured in \[T\] and proved in \[Y\].
A similar enumeration can be done for arbitrary rational knots of given (odd) determinant $`n`$, and one obtains
###### Proposition 5.3
The number of rational knots of determinant $`n`$ ($`n>1`$ odd), counting chiral pairs once, is
$$\frac{1}{4}\left\{\varphi (n)+r_2^0(n)+2^{\omega (n)}\right\},$$
(14)
with $`r_2^0(n)`$ being as in (6), $`\omega (n)`$ denoting the number of different prime divisors of $`n`$ and $`\varphi (n)`$ being Eulerโs totient function.
Proof . We apply Burnsideโs lemma on the action of $`_2\times _2`$ on $`_n^{}`$ given by additive inversion in the first component and multiplicative inversion in the second one. In (14), the second and third term in the braced expression come from counting the square roots of $`1`$ in $`_n^{}`$. These numbers follow from the structure of this group $`_n^{}`$, as remarked in the proof of proposition 4.3. $`\mathrm{}`$
###### Remark 5.1
The functions $`\omega (n)`$ and $`\varphi (n)`$ are hard to calculate for sufficiently large numbers $`n`$ by virtue of requiring the prime factorization of $`n`$, but the expression in terms of these classical number theoretical functions should be at least of theoretical interest.
Counting chiral pairs twice one has the somewhat simpler expression
$$\frac{1}{2}\left\{\varphi (n)+2^{\omega (n)}\right\}.$$
In a similar way one could attempt the enumeration by $`c_p`$ of unknotting number one rational knots of determinant $`p`$ using \[KM\], seeking again an expression in terms of classical number theoretical functions. Obviously from the result of Kanenobu-Murakami we have
$$c_p\mathrm{\hspace{0.17em}2}^{\omega ((p+1)/2)1}+2^{\omega ((p1)/2)1}1,$$
with the powers of two counting the representations of $`(p\pm 1)/2`$ as the product of two coprime numbers $`n_\pm `$ and $`m_\pm `$ up to interchange of factors and the final โ$`1`$โ accounting for the double representation of the twist knot for $`m_+=m_{}=1`$. However, to obtain an exact formula, one encounters the problem that, beside the twist knot, some other knot may arise from different representations (although this does not occur often and the inequality above is very often sharp). For example, for $`p=985`$ the knot $`S(985,288)=S(985,697)`$ occurs for the representations $`m_+=29`$ and $`m_{}=12`$. D. Zagier informed me that he has obtained a complete description of the duplications of the Kanenobu-Murakami forms when considering $`q`$ in $`S(p,q)`$ only up to additive inversion in $`_p^{}`$. According to him, however, considering the (more relevant) multiplicative inversion renders the picture too complicated and number theoretically unilluminating.
## 6 Spanning trees in planar graphs and checkerboard colorings
Here we discuss an interpretation of our results of ยง4 in graph theoretic terms.
###### Theorem 6.1
Let $`n`$ be an odd natural number. Then $`n`$ is the number of spanning trees in a planar self-dual graph if and only if $`n`$ is the sum of two squares.
The proof of this theorem relies on the following construction linking graph and knot theory (see e.g. \[Ka\]). Given an alternating knot (or link) diagram $`D`$, we can associate to it its checkerboard graph.
The checkerboard coloring of a link diagram is a map
$$\{\text{ regions of }D\text{ }\}\{\text{ black, white }\}$$
s.t. regions sharing an edge are always mapped to different colors. (A region is called a connected component of the complement of the plane curve of $`D`$, and an edge a part of the plane curve of the diagram between two crossings.)
The checkerboard graph of $`D`$ is defined to have vertices corresponding to black regions in the checkerboard coloring of $`D`$, and an edge for each crossing $`p`$ of $`D`$ connecting the two black regions opposite at crossing $`p`$ (so multiple edges between two vertices are allowed).
This construction defines a bijection
$$\{\text{ alternating diagrams up to mirroring }\}\{\text{ planar graphs up to duality }\}.$$
Duality of the planar graph corresponds to switching colors in the checkerboard coloring and has the effect of mirroring the alternating diagram if we fix the sign of the crossings so that each crossing looks like rather than . Then we have
###### Lemma 6.1
$`det(D)`$ is the number of spanning trees in a checkerboard graph of $`D`$ for any alternating link diagram $`D`$.
Proof . By the Kauffman bracket definition of the Jones polynomial $`V`$, for an alternating diagram $`D`$, the determinant $`det(D)=|\mathrm{\Delta }_D(1)|=|V_D(1)|`$ can be calculated as follows (see \[Kr\]).
Consider $`\widehat{D}^2`$, the (image of) the associated immersed plane curve(s). For each crossing (self-intersection) of $`\widehat{D}`$ there are 2 ways to splice it:
$$\text{ }\text{}\text{ }\text{ }\text{}\text{ }\text{ or }\text{ }\text{}\text{ }.$$
We call a choice of splicing for each crossing a state. Then $`det(D)`$ is equal to the number of states so that the resulting collection of disjoint circles has only one component (a single circle). We call such states monocyclic.
Let $`\mathrm{\Gamma }`$ be a spanning tree of the checkerboard graph $`G`$ of $`D`$. Define a state $`S(\mathrm{\Gamma })`$ as follows: for any edge $`v`$ in $`G`$ set
$$\text{ }\text{}\text{ }\{\begin{array}{cc}\text{ }\text{}\text{ }& v\mathrm{\Gamma }\\ \text{ }\text{}\text{ }& v\mathrm{\Gamma }\end{array}.$$
Then $`S`$ gives a bijection between monocyclic states of $`D`$ and spanning trees of $`G`$. $`\mathrm{}`$
Since $`det(K)=|\mathrm{\Delta }_K(1)|`$ and it is known that for an $`n`$-component link $`K`$, $`(t^{1/2}t^{1/2})^{n1}\mathrm{\Delta }_K(t)`$, we have that $`2^{n1}det(K)`$. Thus $`det(K)`$ is odd only if $`K`$ is a knot. The converse is also true, since for a knot $`K`$ we have $`\mathrm{\Delta }_K(1)=1`$, and $`\mathrm{\Delta }_K(1)\mathrm{\Delta }_K(1)mod2`$.
Proof of theorem 6.1. If $`n`$ is the number of spanning trees of a planar self-dual graph $`G`$, then its associated alternating diagram $`D`$ is isotopic by $`S^2`$-moves to its mirror image. Since $`n`$ is assumed odd, $`D`$ is a knot diagram. Thus the number of spanning trees of $`G`$, which by the lemma is equal to $`det(D)`$, is of the form $`a^2+b^2`$ by theorem 4.1.
Contrarily, assume that $`n=a^2+b^2`$. Take the checkerboard graph $`G`$ of the diagram in (9) constructed in the proof of theorem 4.2. This diagram has the property of being isotopic to its mirror image by $`S^2`$-moves only (and no flypes), so that its (self-dual) checkerboard graph $`G`$ is the one we sought. $`\mathrm{}`$
Using checkerboard colorings, we will now give a proof of proposition 4.2. We introduce first some more standard notations.
###### Definition 6.1
Given a diagram $`D`$ and a closed curve $`\gamma `$ intersecting $`D`$ in exactly four points, $`\gamma `$ defines a tangle decomposition of $`D`$.
$$D=\text{ }\text{}\text{ }$$
A mutation of $`D`$ is obtained by removing one of the tangles in some tangle decomposition of $`D`$ and replacing it by a rotated version of it by $`180^{}`$ along the axis vertical to the projection plane, or horizontal or vertical in the projection plane. For example:
(To make the orientations compatible, eventually the orientation of either $`P`$ or $`Q`$ must be altered.) $`\gamma `$ is called the Conway circle for this mutation.
Note that a flype can be realized as a sequence of mutations.
There is an evident bijection between the crossings of a diagram before and after applying a mutation, so that we can trace a crossing in a sequence of mutations and identify it with its image in the transformed diagram when convenient. In particular, we can do so for a sequence of flypes.
Proof of proposition 4.2. We proceed by induction on $`n`$. For $`n1`$ the claim is trivial.
Assume now $`L`$ have $`2n`$ crossings and be alternating and achiral, and $`n>1`$. By \[MT\], there is a sequence of flypes (and $`S^2`$-moves) taking an alternating diagram $`D`$ of $`L`$ into its mirror image $`!D`$.
Fix a crossing $`p`$ in $`D`$ and let $`p^{}`$ be the crossing in $`D`$ whose trace under the flypes taking $`D`$ to $`!D`$ takes it to the mirror image of $`p`$ in $`!D`$.
Since the only diagram in which both splicings of a crossing give a nugatory crossing is the Hopf link diagram, for each $`p`$ and $`p^{}`$ there is a splicing not producing a nugatory crossing. We call such splicing a non-nugatory splicing, otherwise call the splicing nugatory.
We can distinguish the two splicings at $`p`$ according to the colors of the regions in the checkerboard coloring they join. We thus call the splicings black or white.
Since whether the black or white splicing gives a nugatory crossing is invariant under flypes, the choice of non-nugatory splicing at $`p`$ and $`p^{}`$ between black or white splicing can be made to be the opposite. Call $`D_p`$ (resp. $`D_p^{}`$) the diagrams obtained from $`D`$ after the so chosen splicing at $`p`$ (resp. $`p^{}`$), and $`D^{}`$ the diagram resulting after performing both splicings. Since $`D_p`$ and $`D_p^{}`$ have no nugatory crossings, $`D^{}`$ is non-split.
We claim that $`D^{}`$ has no nugatory crossings. To see this, use that $`p`$ and $`p^{}`$ join two pairs of regions of opposite color in the checkerboard coloring of $`D`$. If $`D^{}`$ has a nugatory crossing $`q`$, then there would be a closed plane curve $`\gamma `$ intersecting $`D^{}`$ (transversely) only in $`q`$, and lying in some (without loss of generality) white region of the checkerboard coloring of $`D^{}`$.
(15)
But then $`\gamma `$ would persist by undoing the splicing joining the two black regions, and thus $`q`$ would be nugatory in one of $`D_p`$ or $`D_p^{}`$, too, a contradiction.
Since we need the plane curve argument later, let us for convenience call a curve $`\gamma `$ through black (resp. white) regions of $`D`$ intersecting $`D`$ in crossings $`c_1,\mathrm{},c_n`$ a black (resp. white) curve through $`c_1,\mathrm{},c_n`$.
We also claim that $`D^{}`$ depicts the mutant of an achiral link. This follows, because the flypes carrying $`D`$ to $`!D`$ also carry $`D^{}`$ to $`!D^{}`$, modulo mutation. To see this, the only problematic case to consider is when a flype at $`p`$ must be performed. Then
$$\text{ }\text{}\text{ }\text{ }\text{}\text{ }$$
turns into
$$\text{ }\text{}\text{ }\text{ }\text{}\text{ }$$
or
$$\text{ }\text{}\text{ }\text{ }\text{}\text{ },$$
which are both mutations. (Altering the way of building connected sums out of the prime factor diagrams is also considered a mutation.)
Since mutation does not alter the determinant, we have by induction $`det(D^{})(n1)(n4)`$.
Now let $`D_b^{}`$ and $`D_w^{}`$ we the diagrams obtained from $`D`$ when at both $`p`$ and $`p^{}`$ the black resp. white splicings are applied, and $`L_b`$ and $`L_w`$ the alternating links they represent.
When a nugatory white splicing at a crossing $`r`$ in $`D`$ renders a crossing $`q`$ nugatory, then we have a white curve $`\gamma `$ as in (15) though $`p`$ and $`q`$. Similarly if two white splicings at crossings $`r`$ and $`s`$ render $`q`$ nugatory, then we have a white curve $`\gamma `$ in $`D`$ through $`q`$, $`r`$ and $`s`$.
Assume now, a crossing $`q\{p,p^{}\}`$ is nugatory in both $`D_b^{}`$ and $`D_w^{}`$. Then we have a white curve $`\gamma _{q,w}`$ and a black curve $`\gamma _{q,b}`$ in $`D`$ through $`q`$ and at least one of $`p`$ and $`p^{}`$ (and no other crossing different from $`p`$ and $`p^{}`$). By the Jordan curve theorem, $`|\gamma _{q,w}\gamma _{q,b}|=2`$.
If $`\gamma _{q,b}D=\gamma _{q,w}D`$ and $`|\gamma _{q,w}D|=2`$, then $`D`$ is a Hopf link diagram, which we excluded.
Thus $`\{|\gamma _{q,b}D|,|\gamma _{q,w}D|\}=\{2,3\}`$. We claim that for each of the two choices $`|\gamma _{q,b}D|=3`$ and $`|\gamma _{q,w}D|=3`$ there is at most one $`q\{p,p^{}\}`$ satisfying these conditions. This is so, because $`\gamma _{q,b}D=\{q,p,p^{}\}`$ and $`\gamma _{q^{},b}D=\{q^{},p,p^{}\}`$ imply that $`q`$ and $`q^{}`$ have the same pair of opposite black regions. (These are the regions adjacent to exactly one of $`p`$ and $`p^{}`$.) Then every white curve through $`q^{}`$ must pass through $`q`$, contradicting $`\gamma _{q^{},w}D\{q^{},p,p^{}\}`$ if $`qq^{}`$. (For $`|\gamma _{q,w}D|=3`$ switch black and white.)
Therefore, each $`q\{p,p^{}\}`$ is not nugatory at least in one of $`D_b^{}`$ and $`D_w^{}`$, with at most 2 exceptions. Hence, $`c(L_b)+c(L_w)2n4`$, and by Crowellโs result $`det(D_b^{})+det(D_w^{})2n4`$.
Thus finally
$$det(D)det(D^{})+det(D_b^{})+det(D_w^{})(n1)(n4)+2n4=n(n3).$$
(16)
As a graph theoretical application, we obtain:
###### Corollary 6.1
The number of spanning trees in a planar connected self-dual graph with $`2n`$ edges is at least $`n(n3)`$. $`\mathrm{}`$
One can also reformulate theorem 4.4 graph-theoretically. For this one remarks that by \[Me\] an alternating diagram of a composite alternating knot is composite, and checkerboard graphs of such diagrams have a cut vertex. (This is a vertex, which when removed together with all its incident edges, disconnects the graph.) With the same argument as in the proof of theorem 6.1 then one has:
###### Theorem 6.2
Let $`n`$ be an odd perfect square. Then $`n`$ is the number of spanning trees in a planar self-dual graph without cut vertex if and only if $`n1,\mathrm{\hspace{0.17em}9},\mathrm{\hspace{0.17em}49}`$. $`\mathrm{}`$
In the same way one can pose a conjecture which is slightly stronger than conjecture 4.2:
###### Conjecture 6.1
Let $`n`$ be an odd natural number. Then $`n`$ is the number of spanning trees in a planar self-dual graph without cut vertex if and only if $`n`$ is the sum of two squares and $`n1,\mathrm{\hspace{0.17em}9},\mathrm{\hspace{0.17em}49}`$.
## 7 The leading coefficients of the Alexander and HOMFLY polynomial
In this final section we explain another, apparently unrelated, but also very striking occurrence of squares in connection with invariants of achiral knots, namely in the leading coefficients of their Alexander polynomial, and discuss a possible generalization of this property to the HOMFLY, or skein, polynomial. Our results have been obtained independently (but later) by C. Weber and Q. H. Cรขm Vรขn \[VW\].
The skein polynomial $`P`$ is a Laurent polynomial in two variables $`l`$ and $`m`$ of oriented knots and links and can be defined by being $`1`$ on the unknot and the (skein) relation
$$l^1P\left(\text{ }\text{}\text{ }\right)+lP\left(\text{ }\text{}\text{ }\right)=mP\left(\text{ }\text{}\text{ }\right).$$
(16)
This convention uses the variables of \[LM\], but differs from theirs by the interchange of $`l`$ and $`l^1`$.
There is a classic substitution formula (see \[LM\]), expressing the Alexander polynomial $`\mathrm{\Delta }`$, for the normalization so that $`\mathrm{\Delta }(t)=\mathrm{\Delta }(1/t)`$ and $`\mathrm{\Delta }(1)=1`$, as a special case of the HOMFLY polynomial:
$$\mathrm{\Delta }(t)=P(i,i(t^{1/2}t^{1/2})).$$
(17)
We will denote by $`\mathrm{max}\mathrm{deg}_mP`$ the maximal degree of $`m`$ in $`P`$, and by $`\mathrm{max}cf_mP`$ the leading coefficient of $`m`$ in $`P`$. Similarly we will write $`\mathrm{max}\mathrm{deg}\mathrm{\Delta }`$ and $`\mathrm{max}cf\mathrm{\Delta }`$.
### 7.1 Problems and partial solutions
If $`K`$ is an alternating knot, then the HOMFLY polynomial $`P_K[m^2,l^{\pm 2}]`$ is known to be of the form
$$a_{2g}(l)m^{2g}+(\text{lower }m\text{-degree terms}),$$
with $`a_{2g}[l^2,l^2]`$ being a non-zero Laurent polynomial in $`l^2`$ and $`g=g(K)`$ the genus of $`K`$, the minimal genus of an embedded oriented surface $`S^3`$ with $`S=K`$. (See \[Cr\].) If $`K`$ is achiral, then $`a_{2g}(l^2)=a_{2g}(l^2)`$, that is, $`a_{2g}`$ (and, in fact, all the other coefficients of $`m`$ in $`P_K`$) is self-conjugate.
The main problems we consider here can be formulated as follows.
###### Question 7.1
Is $`\mathrm{max}cf\mathrm{\Delta }_K`$ for an achiral knot $`K`$ always a square up to sign, and if $`\mathrm{\Delta }`$ is normalized so that $`\mathrm{\Delta }(t)=\mathrm{\Delta }(1/t)`$ and $`\mathrm{\Delta }(1)=1`$, is the sign $`\text{sgn}(\mathrm{max}cf\mathrm{\Delta }_K)`$ always given by $`(1)^{\mathrm{max}\mathrm{deg}\mathrm{\Delta }_K}`$?
The questions on $`\mathrm{\Delta }`$ can be generalized to $`P`$.
###### Question 7.2
For which large knot classes is it true that achiral knots have $`\mathrm{max}cf_mP`$ of the form $`f(l^2)f(l^2)`$ for some $`f[l]`$?
This is true for several special cases. It appears convenient to compile them into one single statement.
Recall, that a knot $`K`$ is fibered, if $`S^3K`$ fibers over $`S^1`$ (with fiber being a minimal genus Seifert surface for $`K`$), and homogeneous, if it has a diagram $`D`$ containing in each connected component of the complement (in $`^2`$) of the Seifert circles of $`D`$ (called block in \[Cr, ยง1\]) only crossings of the same sign.
###### Proposition 7.1
Let $`K`$ be an ($`+/`$)achiral knot. Then $`\mathrm{max}cf_mP_K`$ is of the form $`f(l^2)f(l^2)`$ for some $`f[x]`$, if
1. $`K`$ is a fibered homogeneous knot,
2. $`K`$ is a homogeneous knot of crossing number at most 16, or
3. $`K`$ is an alternating knot.
From formula (17) it is straightforward that whenever the leading $`m`$-coefficient of $`P_K`$ is of the above form, both the modulus and sign of $`\mathrm{max}cf\mathrm{\Delta }_K`$ are as requested in question 7.1. For these properties we have some more situations where they can be established. We give again the so far complete list of such cases, even if some of them are trivial.
###### Proposition 7.2
Let $`K`$ be an achiral knot. Then $`|\mathrm{max}cf\mathrm{\Delta }_K|`$ is a square, if
1. $`K`$ is a fibered knot,
2. $`K`$ is a knot of crossing number at most 16,
3. $`K`$ is an alternating knot,
4. $`K`$ is strongly achiral, or
5. $`K`$ is negative achiral.
Moreover, $`\text{sgn}(\mathrm{max}cf\mathrm{\Delta }_K)=(1)^{\mathrm{max}\mathrm{deg}\mathrm{\Delta }_K}`$, if
1. $`K`$ is a fibered homogeneous knot,
2. $`K`$ is a knot of crossing number at most 16,
3. $`K`$ is an alternating knot,
4. $`K`$ is strongly achiral, or
5. $`K`$ is negative achiral.
###### Remark 7.1
We omitted to explicitly mention rational knots in proposition 7.2, as we already remarked that they are alternating, and the achiral ones are strongly ($``$)achiral.
In the following subsections we will collect the arguments establishing the conjectured properties in the indicated special cases. Some of them are well-known, or a matter of electronic verification, and thus do not deserve separate proof. These parts are briefly discussed first. Our main result are the statements in the alternating case, which are proved subsequently.
### 7.2 Some known and experimental results
Question 7.1 was the (chronologically) first question I came across, addressing special properties of the leading coefficients of the Alexander and skein polynomial of achiral knots.
This question came up when considering the formula
$$\mathrm{max}cf\mathrm{\Delta }_K=\pm 2^{2g}\underset{i=1}{\overset{2g}{}}a_i$$
for a rational knot $`K=(a_1\mathrm{}a_{2g})`$ with all $`a_i0`$ even. If $`K`$ is achiral, the sequence $`(a_1,\mathrm{},a_{2g})`$ is palindromic, and so we have, up to the sign, the requested property for rational knots. A further larger class of achiral knots satisfying the conjectured condition are the strongly achiral knots (see the Theorem of \[HK\]). Subsequently, I verified all (prime) knots in Thistlethwaiteโs tables \[HT\] up to 16 crossings (note, that the property for a composite knot will follow from that of its factors), and found no counterexample.
Although there seems much evidence for a positive answer to question 7.2, its diagrammatic, and not topological, origin (see ยง7.4) suggests that it may not be true in general, but at least on some (diagrammatically defined) nice knot classes, for example alternating knots.
We now collect the arguments that prove the easier cases of propositions 7.1 and 7.2.
Proof of proposition 7.1 except part 3) and proposition 7.2 except part 3).
alternating knots. As said, we will deal with the squareness and HOMFLY polynomial later. As for the further-going question on the sign, the positive answer follows from the alternation of the coefficients of $`\mathrm{\Delta }`$ proved by Crowell in \[C\], and the property $`\mathrm{\Delta }(1)>0`$ following from Murasugiโs trick, as explained in ยง2.
knots with at most $`16`$ crossings. The answer is also โyesโ for $`16`$ crossing knots. This follows from some experimental results related to question 7.2. It is clear that an answer โyesโ to question 7.2 implies the same answer to question 7.1. This time a computer experiment found that the answer is not positive in general, but the examples showing exceptional behaviour are not quite simple, and required to use the full extent of the tables presently available. Among $`16`$ crossing knots, only three fail to have this property: $`16_{1025717}`$, $`16_{1025725}`$ and $`16_{1371304}`$. They are all $`+`$achiral and have $`P=m^8(l^2+3+l^2)+O(m^6)`$. See figure 4. Since the Alexander polynomial of the three knots has degree 4 and leading coefficient $`1`$, they still conform to the properties requested in question 7.1. Also, at least the first two knots in figure 4 were found to be fibered by the method of \[Ga\], showing that the homogeneity assumption is essential in part 1) of proposition 7.1.
fibered knots. On the other hand, a class of knots where the squareness property of $`\mathrm{max}cf\mathrm{\Delta }`$ is trivial, are the fibered knots, since then $`\mathrm{max}cf\mathrm{\Delta }=\pm 1`$. For fibered homogeneous (in particular, alternating) knots, the other properties also follow easily from known results, because by \[Cr, corollaries 4.3 and 5.3\] and \[MP\] we have for such knots that $`\mathrm{max}cf_mP=l^k`$ for some $`k2`$, and then achirality shows $`k=0`$.
strongly achiral knots. For strongly achiral knots the claims of proposition 7.2 follow directly from the results of \[HK\] stated in theorem 4.5. The only non-obvious property may be the sign of $`\mathrm{max}cf\mathrm{\Delta }`$ for a strongly negative amphicheiral knot. To see this, first normalize the polynomial $`F`$ found by the theorem by some $`+t^n`$ so that $`\mathrm{min}\mathrm{deg}F=\mathrm{max}\mathrm{deg}F`$. Then we have
$$F(t)=\pm F(t^1).$$
(18)
If we normalize $`\mathrm{\Delta }`$ so that $`\mathrm{min}\mathrm{deg}\mathrm{\Delta }=\mathrm{max}\mathrm{deg}\mathrm{\Delta }`$ and $`\mathrm{\Delta }(1)=1`$, then the minimal and maximal degrees show that we must have $`n=0`$ in
$$\mathrm{\Delta }(t^2)=\pm t^nF(t)F(t^1),$$
and the value at $`t=1`$ shows that we must have the positive sign. Denote by $`[X]_i`$ the coefficient of $`t^i`$ in $`X`$. Then, since $`\mathrm{\Delta }(1)=1`$ and $`\mathrm{\Delta }(t)=\mathrm{\Delta }(t^1)`$, the absolute term $`[\mathrm{\Delta }(t)]_0`$ of $`\mathrm{\Delta }(t)`$ is odd. Thus the same is true for $`\mathrm{\Delta }(t^2)=F(t)F(t^1)`$. But
$$\left[F(t)F(t^1)\right]_0=\underset{i=\mathrm{min}\mathrm{deg}F}{\overset{\mathrm{max}\mathrm{deg}F}{}}[F(t)]_i^2,$$
and so from (18) we conclude that $`F`$ must have non-zero absolute term. This determines the sign in (18) to be positive, and then $`\mathrm{max}cfF=\pm \mathrm{min}cfF`$ dependingly on the parity of $`\mathrm{max}\mathrm{deg}F=\mathrm{max}\mathrm{deg}\mathrm{\Delta }`$.
negative amphicheiral knots. Hartley \[Ha\] has extended the result of \[HK\] for strongly negative amphicheiral knots to arbitrary negative amphicheiral knots. Thus the claim follows from the previous argument. $`\mathrm{}`$
###### Remark 7.2
In fact, in \[Kw\], Kawauchi conjectures that the property of the Alexander polynomial of a strongly negative amphicheiral knot he proves with Hartley in \[HK\], and later Hartley \[Ha\] generalizes to an arbitrary negative amphicheiral knot, extends to the Alexander polynomial of an arbitrary amphicheiral knot. This conjecture clearly implies a positive answer to question 7.1. Kawauchiโs conjecture is true in particular for 2-bridge knots, since in \[HK\] he shows that all amphicheiral 2-bridge knots are strongly negative amphicheiral. I verified the conjecture for all prime amphicheiral knots of $`16`$ crossings. Note that Hartley also obtains a condition for positive amphicheiral knots, but it is too weak to address any of our questions.
###### Remark 7.3
Fibered homogeneous knots contain the homogeneous braid knots of \[S\], but also many more. For example, there are 15 fibered homogeneous prime 10 crossing knots, among them 12 alternating and 2 positive ones, which can be shown by the work of \[Cr\] and an easy computer check not to have homogeneous braid representations: $`10_{60}`$, $`10_{69}`$, $`10_{73}`$, $`10_{75}`$, $`10_{78}`$, $`10_{81}`$, $`10_{89}`$, $`10_{96}`$, $`10_{105}`$, $`10_{107}`$, $`10_{110}`$, $`10_{115}`$, $`10_{154}`$, $`10_{156}`$ and $`10_{161}`$.
### 7.3 Two general statements
The remaining cases of propositions 7.2 and 7.1 are included in two more general theorems. The first one generalizes the result of \[MP2\], where it was shown that for an alternating amphicheiral knot the leading coefficient of the Alexander polynomial is not a prime.
###### Theorem 7.1
Let a knot $`K`$ have a homogeneous diagram $`D`$ which can be turned into its mirror image (possibly with opposite orientation) by a sequence of mutations and $`S^2`$-moves (changes of the unbounded region). Then $`|\mathrm{max}cf\mathrm{\Delta }_K|`$ is a square.
Proof of theorem 7.1. Let $`\stackrel{~}{D}`$ denote the mutation equivalence class of a knot diagram $`D`$ (that is, the set of all diagrams that can be obtained from $`D`$ by a sequence of mutations). Assume $`D`$ to be non-split (the split case easily reduces to the non-split one). We consider orientation reversal as a special type of mutation, so a diagram and its inverse belong to the same mutation equivalence class.
As in \[Cr, ยง1\], the Seifert picture of $`D`$ defines a decomposition of $`D`$ into the $``$โproduct (or Murasugi sum) of special alternating diagrams $`D_1,\mathrm{},D_n`$, called blocks. These diagrams may not be prime. Let $`D_{i,1},\mathrm{},D_{i,n_i}`$ be the prime components of $`D_i`$. Note that all $`D_{i,j}`$ are positive or negative (dependingly on $`D_i`$). They will have no nugatory crossings if $`D`$ has neither.
Define
$$(D):=\{\stackrel{~}{D}_{i,j}\}_{i=1,\mathrm{},n,j=1,\mathrm{},n_i}.$$
Here a set is to be understood with the order of its elements ignored, but with their multiplicity counted (i.e., $`\{1,1,2,3\}=\{1,1,3,2\}\{1,2,3\}`$).
Now apply a mutation on $`D`$. The Seifert picture separates the Conway circle into 3 parts $`A`$, $`B`$ and $`C`$.
Because the Conway circle intersects the Seifert picture only in 4 points, all parts $`A`$, $`B`$ and $`C`$ represent connected components of the blocks in $`D`$ they belong to (or possibly connected sums of several such connected components).
Mutation then has the effect of applying mutation on $`B`$ and interchanging and/or reversing $`A`$ and $`C`$. Therefore, $`(D)=(D^{})`$ for any iterated mutant diagram $`D^{}`$ of $`D`$.
If $`D`$ has the property assumed in the theorem, then $`(D)=(!D)`$, or
$$\{\stackrel{~}{D}_{i,j}\}_{i=1,\mathrm{},n,j=1,\mathrm{},n_i}=\{\stackrel{~}{!D_{i,j}}\}_{i=1,\mathrm{},n,j=1,\mathrm{},n_i}.$$
Let $`\varphi :\{\stackrel{~}{D}_{i,j}\}\{\stackrel{~}{D}_{i,j}\}`$ be the bijection induced by $`\stackrel{~}{D}_{i,j}\stackrel{~}{!D_{i,j}}`$.
Since mutation preserves the writhe, $`\varphi `$ has no fixpoints (unless some $`D_{i,j}`$ has no crossings, in which case $`D`$ is split). Thus $`\varphi `$ descends to a bijection
$$\varphi :\{\stackrel{~}{D}_{i,j}:D_{i,j}\text{ positive}\}\{\stackrel{~}{D}_{i,j}:D_{i,j}\text{ negative}\}.$$
Then by \[Mu2\], $`\mathrm{max}cf\mathrm{\Delta }`$ is multiplicative under $``$-product, and hence
$`\mathrm{max}cf\mathrm{\Delta }_D`$ $`=`$ $`{\displaystyle \underset{i,j}{}}\mathrm{max}cf\mathrm{\Delta }_{D_{i,j}}`$
$`=`$ $`{\displaystyle \underset{i,j:D_{i,j}\text{ positive}}{}}\mathrm{max}cf\mathrm{\Delta }_{D_{i,j}}\mathrm{max}cf\mathrm{\Delta }_{!D_{i,j}}`$
$`=`$ $`{\displaystyle \underset{i,j:D_{i,j}\text{ positive}}{}}\mathrm{max}cf\mathrm{\Delta }_{D_{i,j}}\pm \mathrm{max}cf\mathrm{\Delta }_{D_{i,j}}`$
$`=`$ $`\pm \left({\displaystyle \underset{i,j:D_{i,j}\text{ positive}}{}}\mathrm{max}cf\mathrm{\Delta }_{D_{i,j}}\right)^2,`$
as desired. $`\mathrm{}`$
###### Theorem 7.2
Under the same assumption as theorem 7.1 we have $`\mathrm{max}cf_mP(D)=f(l^2)f(l^2)`$ for some $`f[x]`$.
Proof . Using \[MP2\] instead of \[Mu2\], we obtain $`\mathrm{max}cf_mP(D)=f(l)f(l^1)`$. Since $`\mathrm{max}cf_mP(D)`$ has only even powers of $`l`$, the result follows. $`\mathrm{}`$
###### Corollary 7.1
For any alternating achiral knot $`K`$, $`|\mathrm{max}cf\mathrm{\Delta }_K|`$ is a square and $`\mathrm{max}cf_mP_K=f(l^2)f(l^2)`$.
Proof . Use that an alternating diagram is homogeneous, \[MT\], and that a flype can be realized as a sequence of mutations. $`\mathrm{}`$
Note, that in the case of $`\mathrm{\Delta }`$ we obtain an exact condition when a number occurs as $`|\mathrm{max}cf\mathrm{\Delta }_K|`$ for an alternating achiral knot $`K`$ (since the other implication is trivial).
###### Remark 7.4
We need the homogeneity of $`D`$ only to assure that all $`D_{i,j}`$ are positive or negative. (This weaker property is invariant under mutations, whereas homogeneity is not.) Thus the theorems could be formulated even slightly more generally, but then also more technically.
Although the corollary is the most interesting special case of the theorems, they give indeed more general statements.
###### Example 7.1
The non-alternating achiral knots $`14_{45317}`$ and $`14_{45601}`$ have unique minimal diagrams (which therefore must be transformable into their mirror images by $`S^2`$-moves only), which are both homogeneous (of genus $`5`$ and $`4`$, respectively).
###### Remark 7.5
Here we consider (and mean unique) minimal diagrams only up to $`S^2`$-moves, and not as in \[HT\] up to $`S^2`$-moves and mirroring. There are achiral knots with a unique minimal diagram up to $`S^2`$-moves and mirroring, but corresponding to two different (mirrored) diagrams up to $`S^2`$-moves only, which are not interconvertible by flypes. One such example is $`14_{41330}`$.
###### Remark 7.6
If a knot has a diagram $`D`$, which can be transformed into its (possibly reverted) obverse by moves in $`S^2`$ and flypes, then it also has a diagram $`D^{}`$, which can be transformed into its obverse by moves in $`S^2`$ only. This follows from the fact that mirroring and moves in $`S^2`$ take the flyping circuits of $`D`$ into each other (for the definition of flyping circuits see \[ST, ยง3\]), and flyping in a flyping circuit is independent from the other ones. $`D^{}`$ can be obtained by appropriate flypes from $`D`$. (For an analogous statement about the checkerboard graphs, see \[DH\].)
### 7.4 Some diagrammatic questions
The two theorems of ยง7.3 suggest the diagrammatic arguments motivating questions 7.1 and 7.2. We conclude with a more detailed problem concerning possible generalizations. In order to make the result of \[MP\] work, we need to consider $`P`$-maximal diagrams.
###### Definition 7.1
Call a link diagram $`D`$ with $`c(D)`$ crossings and $`s(D)`$ Seifert circles $`P`$-maximal, if $`\mathrm{max}\mathrm{deg}_mP(D)=c(D)s(D)+1`$.
In \[Mo\], Morton showed that the one inequality $`\mathrm{max}\mathrm{deg}_mP(D)c(D)s(D)+1`$ holds for any arbitrary link diagram, and used this to show that there are knots $`K`$, which do not possess a diagram $`D`$ with $`g(D)=g(K)`$ (a fact that also implicitly follows from \[Wh\]). Here $`g`$ denotes the genus of a knot or knot diagram, for latter being defined by $`g(D)=1/2(c(D)s(D)+1)`$, which is the genus of the canonical Seifert surface associated to $`D`$ (see \[Ad, ยง4.3\] or \[Ro\]).
In \[Cr\] it was shown that homogeneous diagrams are $`P`$-maximal. Many knots have $`P`$-maximal diagrams โ beside the homogeneous knots, for example all (other) knots in Rolfsenโs tables \[Ro, appendix\] and also all the 11 and 12 crossing knots tabulated in \[HT\]. However, some knots do not โ in \[St2, fig. 9\] we gave four examples of 15 crossings.
In \[MP\] it was shown that $`\mathrm{max}cf_mP`$ is multiplicative under $``$-product of $`P`$-maximal diagrams. This is the link between the above polynomial conjectures and the diagrammatic problems, which we summarize in the question below.
###### Question 7.3
Does any ($`+/`$)achiral knot (or an achiral knot in which large knot class) have a diagram that can be
1. transformed into its (possibly reverted) obverse by moves in $`S^2`$ (changes of the unbounded region), or
2. represented as the iterated connected and Murasugi sum of (some even number of) $`P`$-maximal link diagrams $`D_i`$ with $`\{D_i\}`$ being mutually obverse (up to orientation and $`S^2`$-moves) in pairs?
As a motivation, we mention briefly some related partial cases and implications.
Remarks on part A):
* By considering the blocks of such a diagram, we see that it is the $``$-product of special diagrams $`D_i`$, such that each $`D_i`$ is transformable by $`S^2`$-moves into the obverse of itself, or of some other $`D_j`$, $`ji`$.
* If the answer is โyesโ for some homogeneous diagram (which in particular happens by \[MT\] for alternating diagrams not admitting flypes), no blocks transform onto their own obverses, and \[MP\] shows a positive answer to question 7.2.
The $`(l)`$coefficients of $`\mathrm{max}cf_mP`$ of the three knots in figure 4 do not alternate in sign, and so these knots cannot be homogeneous (beware of the different convention for $`P`$ in \[Cr\]!). Still, the first two knots have a $`P`$-maximal diagram of the type requested in part A). It can be obtained from the one given in the figure by a flype. Thus theorem 7.2 does not hold under the weaker assumption the diagram to be $`P`$-maximal instead of homogeneous.
* The answer is โyesโ for rational knots. The (palindromic) expression $`(a_1\mathrm{}a_na_n\mathrm{}a_1)`$ with all $`2g`$ even numbers $`a_i0`$ gives a rational diagram (of the form $`D!D`$, where $`D`$ is a connected sum of diagrams of reversely oriented $`(2,a_{2i})`$-torus links), having the desired property.
* By remark 7.6, one can equivalently also allow moves in $`S^2`$ and flypes, so the answer is in particular positive for alternating knots by \[MT\].
Remarks on part B):
* By \[MP\], a positive answer implies a positive answer also for question 7.2 (thus in particular the answer is negative for the knots on figure 4). This is the motivation for proposing question 7.2 after arriving to question 7.1.
* In turn, a positive answer to B) is implied by a positive answer to A) for homogeneous diagrams. This is the motivation for proposing part B).
Acknowledgements. I would like to thank to Kunio Murasugi, Morwen Thistlethwaite, Kenneth Williams, and especially to Don Zagier for their helpful remarks. |
warning/0003/hep-ph0003155.html | ar5iv | text | # NEW PHYSICS SEARCH IN B MESON DECAYS
## 1 Introduction
$`B`$ fatories at KEK and SLAC are taking data to probe the origin of CP violation which is one of main issue in current high energy physics. In the standard model(SM) the CP violation is originated by a physical phase of the Cabbibo-Kobayashi-Maskawa(CKM) matrix $`^\mathrm{?}`$. A new source of CP violation can appear in models beyond the SM. If there is new physics beyond SM, we expect to see its effects in CP violating B meson decays.
The most important task for new physics search is to identify decay modes where one can find a large deviations from standard model expectations, and also experimentally accessible in near future. It is really good chance to probe the new physics effects in B-meson decays through indirect ($`B^0\overline{B}^0`$ mixing) and direct CP-violating processes at B factories.
In this talk we discuss a possible new physics impact on B-meson rare decays and the direct CP violation phenomena through the magetic gluon penguin contributions in $`B^\pm K^\pm \varphi ,K^0\pi ^\pm `$ decays.
## 2 New physics Effects in mSUGRA Model
We investigate the new physics effects in rare B decays: $`BX_s\gamma `$ and $`BX_s\mathrm{}^+\mathrm{}^{}`$ and in direct CP violating modes: $`B^\pm \varphi K^\pm ,K^0\pi ^\pm `$ within the minimal Supergravity model(mSUGRA). In the mSUGRA, there are four new CP violating phases, i.e. phases of the gaugino mass, the higgsino mass parameter, the SUSY breaking Higgs beson mass, and the trilinear scalar coupling constant, of which two combinations are physically independent. When we impose the universal condition at GUT scale, two physical phases at GUT scale comes from, if we take a phase convention, the trilinear coupling constant and higgsino mass parameter, $`\varphi _A`$ and $`\varphi _\mu `$, respectively. These phases induce the neutron and electron electric dipole moments (EDMs). When we require the unversality of SUSY breaking term at GUT scale and explicitly solve the renormalization group equations (RGEs) to determine the masses and the mixings of SUSY particles and also require the condition for the radiative electroweak breaking, the phase $`\varphi _\mu `$ is strongly constrained by EDMs and the phase $`\varphi _A`$ is not constrainted at GUT scale, however, in low energy scale, the phase of A-term for top squarks is strongly supressed, becuase the phase of A-term for top squarks is reduced due to the large top Yukawa coupling constant and aligned to that of the gaugino mass $`^\mathrm{?}`$.
When we investigate the effect of the SUSY CP violating parameter on rare B decays, $`BX_s\gamma `$ and $`BX_s\mathrm{}^+\mathrm{}^{}`$, in the mSUGRA, we find some interesting results $`^\mathrm{?}`$ with following SUSY parameters: $`0<m_0<1`$ TeV, $`120<M_X<500`$ GeV, $`|A_x|<5m_0`$, and the bound of EDMs$`^\mathrm{?}`$ : $`|d_n|0.97\times 10^{25}ecm`$, $`|d_e|4.0\times 10^{27}ecm`$, in addition, the branching ratio of $`BX_s\gamma `$ $`^\mathrm{?}`$ : $`2.0\times 10^4<(BX_s\gamma )<4.5\times 10^4`$ ; (i) $`\varphi _\mu 10^2`$ and $`0\varphi _A2\pi `$, (ii) As in the case of no SUSY CP violating phase, $`C_7`$ and $`C_8`$ have large SUSY contributions, however, those to $`C_9`$ and $`C_{10}`$ are small. (iii) CP asymmetry of $`BX_s\gamma `$ : $`A_{CP}(BX_s\gamma )2\%`$ with EDM constraints, however when we consider EDM cancellation in one loop level $`^\mathrm{?}`$, it can be reached upto $`7\%`$. (iv) In $`BX_s\mathrm{}^+\mathrm{}^{}`$ decay, for small $`tan\beta `$ value, $`Im(C_7/C_7^{sm})0`$ since $`Im(A_t)`$ becomes small, however, for large $`tan\beta `$ value, since chargino and stop loop effect becomes dominated in $`C_7`$, $`C_7\pm C_7^{sm}`$. So branching ratio of $`BX_s\mathrm{}^+\mathrm{}^{}`$ can be enchanced when $`C_7C_7^{sm}`$. (v) The allowed domain of $`C_8/C_8^{sm}`$ can be extracted from the $`BX_sg\gamma `$ contribution into $`BX_s\gamma `$, shown in Figure 1.
When we investigate the new physics effect through magnetic gluon penguin contributions in the exclusive B meson decays, we find that a possible large direct CP violation can be observed in pure penguin modes, specially $`B^\pm \varphi K^\pm `$ and $`K^0\pi ^\pm `$.
The CP asymmetry is defined as :
$`A_{CP}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(B^{}f)\mathrm{\Gamma }(B^+\overline{f})}{\mathrm{\Gamma }(B^{}f)+\mathrm{\Gamma }(B^+\overline{f})}}`$
For instance, the amplitude of $`B^+\varphi K^+`$ decay is :
$`A(B^+\varphi K^+)`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}\left[a_3+a_4+a_5{\displaystyle \frac{1}{2}}(a_7+a_9+a_{10})+a_{8G}\right]M^{(BK,\varphi )}`$
$`M^{(BK,\varphi )}`$ $`=`$ $`<\varphi |(\overline{s}s)_{VA}|0><K|(\overline{b}s)_{VA}|B>=2f_\varphi m_\varphi [ฯตP_B]F_1^{BK}(m_\varphi ^2)`$
where
$`a_{8G}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}{\displaystyle \frac{m_b^2}{q^2}}C_{8G}{\displaystyle \frac{N_c^21}{N_c^2}}S_{\varphi K}e^{i\sigma }`$
$`C_{8G}`$ $`=`$ $`C_{8G}^{new}+C_{8G}^{sm}=C_{8G}^{sm}Re^{i\theta }`$
where $`S_{\varphi K}=0.49`$, $`\sigma `$ is the strong phase difference between $`O_{8G}`$ and $`O_{110}`$, $`\theta =\delta _{new}\delta _{sm}`$ is the electroweak phase difference between new physics and SM, and $`R=|C_{8G}/C_{8G}^{sm}|`$. In our analysis we use the factroization approach including non-factorizable contributions into $`N_c^{eff}`$ and strong phases via Bander-Silverman-Soni mechanism $`^\mathrm{?}`$ into $`a_i`$ and $`C_i^{eff}`$. We use $`(N_c^{eff})_{LL}=2.0`$ for $`O_{1,2,3,4,9,10}`$ and $`(N_c^{eff})_{LR}=6.0`$ for $`O_{5,6,7,8}`$ as like as H.Y. Cheng et al.$`^\mathrm{?}`$. As shown in Figure 2, we have large direct CP asymmetries which is induced by new physics contributions : in $`B^\pm \varphi K^\pm `$ decay, we have $`0.1\times 10^5(B^\pm \varphi K^\pm )0.75\times 10^5`$ and $`|A_{CP}|15\%`$ with EDM constrained condition, however, without EDM constraints, $`|A_{CP}|30\%`$. For $`B^\pm K^0\pi ^\pm `$ decay, we get the branching ratio upto $`22.5\times 10^6`$ which is well agreed with present experimental data $`^\mathrm{?}`$: $`(B^\pm K^o\pi ^\pm )=(18.2_{4.0}^{+4.6}\pm 1.6)\times 10^6`$ and $`|A_{CP}|20\%`$.
In conclusion we have given example of decay modes which can allow an early detection of new physics effects in the minimal supergravity model.
## Acknowledgments
The author wishes to thank T. Goto, T.Nihei, Y. Okada and Y.Shimizu for collaboration on some of the work presented here. Y.Y.K. would like to thank M. Kobayashi and H.Y Cheng for their hospitality.
## References |
warning/0003/astro-ph0003403.html | ar5iv | text | # The Anisotropy of MHD Alfvรฉnic Turbulence
## 1 Introduction
Many astrophysical plasmas, including the interstellar medium and the solar wind, often show magnetic fields whose energy density is greater than or equal to the local kinetic energy density. In these plasmas the magnetic fields play a dominant dynamical role, mediated by magnetohydrodynamic (MHD) waves. In the incompressible limit, there are only two types of linear modes: shear Alfvรฉn waves and pseudo Alfvรฉn waves. While these two modes have different polarization directions, they have the same dispersion relation and propagate along the magnetic field lines at the Alfvรฉn speed. Therefore, the nonlinear interactions of wave packets moving along the magnetic field lines at the Alfvรฉn speed determine the dynamics of incompressible magnetized plasmas with a strong background field. In this paper, we study the anisotropy of the MHD turbulence in this regime. We will refer to this turbulence as incompressible Alfvรฉnic turbulence.
Nonlinear processes and the corresponding energy spectrum of incompressible Alfvรฉnic turbulence are still among the most controversial problems in MHD. Since the pioneering works of Iroshnikov (1963) and Kraichnan (1965), the Iroshnikov-Kraichnan (IK) theory has been widely accepted as a model for incompressible, highly conducting MHD turbulence. The IK theory predicts $`E_M(k)E_K(k)k^{3/2}`$ from a Kolmogorov-like dimensional analysis. Here, $`E_M(k)`$ and $`E_K(k)`$ are the magnetic and kinetic energy spectra respectively. In this framework, two counter-traveling eddies (i.e. Alfvรฉn wave packets) interact and transfer energy to smaller spatial scales only when they collide, as they move in opposite directions along the magnetic field lines. Since the duration of such a collision is shorter than the conventional eddy turnover time by a factor of $`t_v(l)/t_A(l)`$, this collisional process is inefficient and the spectral energy transfer time as a function of scale $`l`$ ($`=t_{cas}(l)`$) increases by the same factor compared to the eddy turnover time ($`l/v_l`$) in ordinary hydrodynamic turbulence. Here $`t_v(l)=l/v_l`$ and $`t_A(l)=l/V_A`$ are eddy turnover time and Alfvรฉn time respectively, $`V_AB/\sqrt{4\pi \rho }`$, and $`B`$ is the rms magnetic field strength. When the external field is strong, as assumed in IK theory, this quantity is usually set to $`B_0`$, the strength of the uniform background field. If the spectral energy cascade rate
$$ฯต\frac{v_l^2}{t_{cas}(l)}\frac{v_l^3}{l}\frac{t_A(l)}{t_v(l)}$$
(1)
is scale-independent and $`E_M(k)E_K(k)`$, then we obtain the IK energy spectra.
The IK theory assumes an isotropic distribution of energy in $`๐ค`$-space. However, many researchers have argued that anisotropy is an important characteristic in MHD turbulence (for example, Shebalin et al 1983, Montgomery and Matthaeus 1995). This anisotropy results from the resonant conditions for 3-wave interactions (or 4-wave interactions, when 3-wave interactions are null). The resonant conditions for the 3-wave interactions are
$`๐ค_1+๐ค_2`$ $`=`$ $`๐ค_3,`$ (2)
$`\omega _1+\omega _2`$ $`=`$ $`\omega _3,`$ (3)
where $`๐ค`$โs are wavevectors and $`\omega `$โs are wave frequencies. The first condition can be viewed as momentum conservation and the second as energy conservation. Alfvรฉn waves satisfy the dispersion relation: $`\omega =V_A|k_{}|`$, where $`k_{}`$ is the component of wavevector parallel to the background magnetic field. Since only opposite-traveling wave packets interact, $`๐ค_1`$ and $`๐ค_2`$ must have opposite signs. Then from equations (2) and (3), either $`k_{,1}`$ or $`k_{,2}`$ must be equal to 0. That is, zero frequency modes are essential for energy transfer. If $`k_{,2}=0`$, we have
$`k_{,1}`$ $`=`$ $`k_{,3},`$ (4)
$`k_{,2}`$ $`=`$ $`0`$ (5)
(Shebalin et al 1983). Therefore, in the wavevector space, 3-wave interactions make energy cascade in directions perpendicular to the mean magnetic field. Since the energy cascade is strictly perpendicular to the mean magnetic field, the actives modes in wavevector space have a slab-like geometry with a constant width. The implication is that the nonlinear cascade of energy works against isotropy in $`๐ค`$ space. Furthermore, it is important to note that equations (2) and (3) are true only when wave amplitudes are constant. In reality, nonlinear interactions provide a natural broadening mechanism, following the uncertainty relation, $`\mathrm{\Delta }t\mathrm{\Delta }\omega 1`$. In particular, if a wave has a frequency less than or comparable to the nonlinear interaction rate, it is effectively a zero frequency mode.
Goldreich and Sridhar (1995, 1997) showed that in the strong incompressible shear Alfvรฉnic turbulence regime (i.e. $`\tau _{NL}^1\omega `$), these arguments lead to a new scaling law with a scale-dependent anisotropy. In this model smaller eddies are more elongated. Their arguments are based on the assumption of a critically balanced cascade, $`k_{}V_Ak_{}v_l`$, where $`k_{}`$ and $`k_{}`$ are wave numbers perpendicular and parallel to the external dc field<sup>1</sup><sup>1</sup>1Later in this paper, we will use $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$, instead of $`k_{}`$ and $`k_{}`$, to represent their scaling relation in a slightly different, yet we believe equivalent, viewpoint. Here, $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$ are wavenumbers with respect to the direction of the local magnetic field, not the external field.. The argument given above for 3-wave interactions makes it clear that $`k_{}`$ will tend to increase until it becomes important in the plasma dynamics. The assumption of strong nonlinearity implies that wave packets lose their identity after they travel one wavelength along the field lines. Consequently the eddy turnover time ($`(k_{}v_l)^1`$) is actually the same as Alfvรฉnic time ($`(k_{}V_A)^1`$). In this model, the cascade time, $`t_{cas}(l)`$ can be determined without ambiguity: $`t_{cas}(k_{}v_l)^1(k_{}V_A)^1`$. Since the cascade time is comparable to the period of the Alfvรฉn wave, the 3-wave resonant condition can be violated according to the uncertainty relation $`\mathrm{\Delta }\omega \mathrm{\Delta }tV_Ak_{}t_{cas}1`$. The quantity $`k_{}`$ is the width of the active region in wavevector space. Finally the assumption of a scale-independent cascade rate $`ฯตv_l^2/t_{cas}(l)E_{waves}V_A/L`$ gives
$`k_{}k_{}^{2/3}L^{1/3}\left({\displaystyle \frac{E_{waves}}{V_A^2}}\right)^{1/3},`$ (6)
$`v_lV_A(k_{}L)^{1/3}\left({\displaystyle \frac{E_{waves}}{V_A^2}}\right)^{1/3},`$ (7)
where $`E_{waves}`$ is the wave energy per mass. These formulae assume that all scales, from $`L`$ on down, are within the inertial range of MHD turbulence. Here the typical $`k_{}`$ should be interpreted as the size of the range of parallel wavevectors, corresponding to a given $`k_{}`$, that contain significant energy.
Matthaeus et al. (1998) recently tested this model numerically, and showed that the anisotropy of low frequency MHD turbulence scales linearly with the ratio of perturbed and total magnetic field strength ($`b/B`$), a result which seems inconsistent with Goldreich and Sridharโs model. To explain this scaling relation, they suggested that the region of Fourier space where the energy transfer takes place actively is given by
$$|๐ค\frac{๐_0}{\sqrt{4\pi \rho }}|<\frac{1}{\tau _{NL}},$$
(8)
where $`\tau _{NL}`$ is the eddy turnover time of the energy containing length $`L`$. Consequently, โthe region of the wave number space where spectral transfer is most vigorousโ has a slab-like geometry with a constant width proportional to $`1/(\tau _{NL}B_0)`$.
All these theories (except the IK theory) share a common prediction for anisotropy: anisotropy should be more pronounced on smaller scales. Oughton et al. (1994) and Ghosh and Goldstein (1997) already reported this scale-dependent anisotropy through numerical simulations. The former extended Shebalin et al. (1983)โs 2-D calculations to 3-D cases. To measure anisotropy, they introduced the Shebalin angles, $`\theta _Q`$, defined by $`\mathrm{tan}^2\theta _Q=(k_{}^2|๐(๐ค,t)|^2)/(k_{}^2|๐(๐ค,t)|^2),`$ where $`๐`$ is vector potential A, magnetic field B, or current J, etc. Greater $`\theta _Q`$ means greater anisotropy. They found that $`\theta _A<\theta _B<\theta _J`$. If the energy spectrum of B scales as $`E_M(k)k^s`$, then the spectra of vector potential and current scale as $`E_A(k)k^{s2}`$ and $`E_J(k)k^{s+2}`$, respectively. The spectra of vector potential has the steepest slope among the three spectra. This means that the vector potential is least strongly dependent on small scales (and the current is most strongly dependent on small scales). Therefore they concluded that anisotropies are more pronounced at smaller scales<sup>2</sup><sup>2</sup>2 The width of the active region in Fourier space, $`k_{}`$, is a function of $`k_{}`$. If
$$\left[\frac{k_{}}{k_{}}\right]_{\text{small }k_{}}>\left[\frac{k_{}}{k_{}}\right]_{\text{large }k_{}},$$
then we will have the ordering of Shebalin angles as observed by Shebalin et al. and Oughton et al. However, if $`k_{}k_{}`$, then we do not expect any ordering among the angles. In ยง3, we will find that, no matter what the true functional form $`k_{}=k_{}(k_{})`$ is, Fourier transformation smooths out the true relation and leads to a fake linear relationship between $`k_{}`$ and $`k_{}`$. This suggests that their results are in contradiction to our discussion in ยง3. However, the apparent linear relationship between $`k_{}`$ and $`k_{}`$ in ยง3 is not perfectly linear. Instead, we expect $`k_{}=c_1k_{}+c_2`$, where $`c_1`$ is a decreasing function of $`B_0`$ and $`c_2`$ depends on the $`k_{}`$ of forcing terms (or initial values of $`k_{}`$ for decaying turbulence). The presence of $`c_2`$ does not seem to be important in our Fig. 3 in ยง3. However, it does affect the calculation of Shebalin angles. That is, because of $`c_2`$, the ratio $`k_{}/k_{}=c_1+c_2/k_{}`$ becomes a function of $`k_{}`$. At the largest energy containing eddy scale, $`k_{}c_2`$ and hence $`k_{}/k_{}c_1+O(1)O(1)`$. But, at small scales, the ratio can be much smaller than unity. Therefore, because of $`c_2`$, we can obtain a scale-dependent anisotropy: $`(k_{}/k_{})`$ at small $`k_{}`$ is greater than that at large $`k_{}`$. Note that this is a result of the initial conditions, rather than a true scaling relation. This will lead the ordering of the angles as observed by Shebalin et al. and Oughton et al.. This interpretation is qualitatively consistant with their results. For example. the ratio $`k_{}/k_{}=c_1+c_2/k_{}^{peak}`$, therefore $`(\mathrm{tan}\theta _Q)^1`$, approaches to a constant value $`c_2/k_{}^{peak}`$ as $`B_0`$ becomes strong. Here $`k_{}^{peak}`$ is the wavenumber of the peak of the energy spectrum. It is also possible to explain the increased anisotropy at high magnetic Reynolds numbers. If the magnetic Reynolds number increases, then $`k_{}^{peak}`$ increases and, therefore, the ratio decreases. . They also found a similar ordering for velocity field (and vorticity, etc). On the other hand, Ghosh and Goldstein (1997) calculated the Shebalin angles as a function of wavenumber bin. They found that the Hall-MHD<sup>3</sup><sup>3</sup>3Hall MHD includes the Hall term, which is important at ion-cyclotron scales. In this paper, we consider only the standard MHD equations. simulations show increased anisotropy at small scales (i.e. greater Shebalin angles at smaller scales). However their standard MHD simulations do not show increased anisotropies at small scales. We refrain from comparing their work with ours because they used different physics (the Hall effect) and their simulations are $`2\frac{1}{2}`$ dimensional. The simulations given in this paper are 3-D standard MHD simulations. We note that none of the previous papers quantitatively compared their results with particular theories of anisotropy.
In this paper, we examine the scaling law for Alfvรฉnic MHD turbulence numerically and resolve the controversy concerning the anisotropic structure of the turbulence. Our results are consistent with Goldreich and Sridharโs model. We describe our numerical methods in ยง2. In ยง3, we describe our results for anisotropy in wavevector space. In this section, we demonstrate that none of the scaling laws mentioned above agrees with our data, and explain why a straightforward evaluation of the distribution of spectral power does not correspond to a physically meaningful set of scaling laws. We attribute this to the systematic effects of large scale curvature of the magnetic fields. In ยง4, we determine the shape of individual eddies, avoiding the systematic error described above. We compare our results to Goldreich and Sridharโs model and give our conclusions in ยง5.
## 2 Numerical Methods and General Results
We have employed a pseudospectral code to solve the incompressible MHD equations in a periodic box of size $`2\pi `$:
$$\frac{๐}{t}=(\times ๐)\times ๐V_{A0}^2(\times ๐)\times ๐+\nu ^2๐+๐+P^{},$$
(9)
$$\frac{๐}{t}=๐๐๐๐+\eta ^2๐,$$
(10)
where $`๐`$ is random driving force and $`P^{}P+๐๐/2`$. Other variables have their usual meaning. $`V_{A0}=B_0/(4\pi \rho )^{1/2}`$ is the Alfvรฉn velocity of the background field, which is set to be order unity in our simulations. In pseudo spectral methods, we calculate the temporal evolution of the equations (9) and (10) in Fourier space. To obtain the Fourier components of nonlinear terms, we first calculate them in real space, and transform back into Fourier space. We use 21 forcing components with $`2k\sqrt{12}`$. Each forcing component has correlation time of one. The peak of energy injection occurs at $`k2.5`$. The amplitudes of the forcing components are tuned to ensure $`V1`$. Since we expect the perturbed magnetic field strength $`b`$ to be comparable to or less than $`V`$ (i.e. $`b1`$), the resulting turbulence can be regarded as strong MHD turbulence (i.e. $`VB_0B`$). The average helicity in these simulations is not zero. However, our results are insensitive to the value of helicity, possibly because the box size is only slightly bigger than the energy injection scale. We use an appropriate projection operator to calculate $`P^{}`$ term in Fourier space and also to enforce divergence-free condition ($`๐=๐=0`$). We use up to $`256^3`$ collocation points. We use integration factor technique for kinetic and magnetic dissipation terms and leap-frog method for nonlinear terms. We eliminate $`2\mathrm{\Delta }t`$ oscillation of the leap-frog method by using an appropriate average. At $`t=0`$, magnetic field has only uniform component and velocity has a support between $`2k4`$ in wavevector space.
We use hyperviscosity and hyperdiffusivity for dissipation terms to maximize the inertial range. The only exception is Run 256P-$`B_0`$1, where we use physical viscosity and diffusivity. The power of hyperviscosity is set to 8, so that the dissipation term in the above equation is replaced with
$$\nu _8(^2)^8๐,$$
(11)
where we set the value of $`\nu _8`$ from the condition $`\nu _h(N/2)^{2h}\mathrm{\Delta }t0.5`$ (see Borue and Orszag 1996). Here $`\mathrm{\Delta }t`$ is the time step and $`N`$ is the number of grids in each direction. We use exactly same expression for the magnetic dissipation term. We list parameters used for the simulations in Table 1. We use the notation 256X-Y, where X = H or P refers to hyper- or physical viscosity; Y = $`B_0`$1 or $`B_0`$0.5 refers to the strength of the external magnetic fields.
Fig. 1 shows the evolution of $`V^2`$ and $`B^2`$ from Runs 256H-$`B_0`$1 and REF2. Results of 256H-$`B_0`$1 are plotted as solid lines, and results of REF2 are plotted as dotted lines. Both runs have similar results for the overlapped time, which means that the values of $`V^2`$ and $`B^2`$ are not sensitive to the spatial resolution. The magnetic energy density grows fast at the beginning of the simulations, as a result of the stretching of external field lines by turbulent motion. Since the external field is strong, magnetic forces soon become strong enough to balance the stretching effect. This balance happens at $`t1`$ and, after a transient stage ($`1t5`$), the turbulence reaches the saturation stage ($`t5`$). In this case the saturation stage begins quite early, since the external uniform magnetic field is strong and the magnetic diffusivity is effectively 0. In general, when the external field is weaker and/or magnetic diffusivity is larger, the saturation stage occurs later. At the saturation stage, $`B^21.45`$ and $`V^20.6`$. Since $`b^2=B^2B_0^2(1)0.45`$, there is a rough energy equipartition between $`b`$ and $`V`$. Note that, since $`VB`$, the condition for strong MHD turbulence is met.
## 3 Fourier Analysis: A False Scaling Law?
### 3.1 Fourier Analysis
In this subsection, we investigate the spectral structure of Alfvรฉnic turbulence. It is important to note that we lose some information about individual eddies when we perform a global transformation, such as the Fourier transformation. In particular, the scaling law given in this subsection may $`not`$ be true for individual eddies (See ยง3.2 for details).
We plot the 1-dimensional energy spectra $`E_K(k)`$ and $`E_M(k)`$ in Fig. 2. Both spectra peak at the same $`k`$, which is also the scale of the energy injection. This reflects that there is a rough energy equipartition between $`b`$ and $`V`$ at the largest energy containing scale. In general, $`E_K(k)`$ always peaks at the energy injection scale. However, $`E_M(k)`$ peaks at a larger k than $`E_K(k)`$ when the external field is weak. The inertial range exhibits of two different scaling ranges: a small $`k`$ range where the slopes are steep and a large $`k`$ range where the slopes are mild. We believe that the former reflects a true inertial range, whereas the latter is a result of the $`bottleneck`$ effect. The $`1/k`$ bottleneck effect is a common feature in numerical hydrodynamic simulations with hyperviscosity (see Borue and Orzag 1996, 1995). Interestingly enough, the slope of the bottleneck is actually steeper than $`1`$. This might mean that the bottleneck effect is less serious in the MHD case. The kinetic and magnetic energy spectra have slightly different powers in the inertial range. Although the power indexes of both spectra are close to $`5/3`$ in the (true) inertial range, the data are also compatible with IK theory, where the power index is $`3/2`$.
We plot normalized 3-dimensional energy spectra of Run 256H-$`B_0`$1 in Fig. 3. In the figure, we plot contours of same
$$E_3(k_{},k_{})/E_3(k_{},0)$$
(12)
in the $`(k_{},k_{})`$ plane. As the energy cascades in the directions perpendicular to the mean field, this ratio drops as we move away from the $`k_{}`$ axis. The figure implies most of the energy is confined in a region around the $`k_{}`$ axis. The thickness of the region depends on the strength of the external magnetic fields (Fig. 4). We use the contours of $`E_3(k_{},k_{})/E_3(k_{},0)=0.5`$ to measure the thickness. The angle $`\mathrm{\Theta }_{0.5}`$ is the angle formed by the contours and the horizontal axis. We can see that $`\mathrm{tan}\mathrm{\Theta }_{0.5}`$ ($`=k_{}/k_{}`$) is proportional to $`b/B_0`$. This result confirms the scaling relation found by Matthaeus et al. (1998). Both velocity and magnetic fields have very similar structures. From the fact that $`k_{}<k_{}`$ in the active region (see Fig. 3), one can conclude that eddies do have anisotropic structure: eddies are stretched along the direction of the mean field. Apparently, Fig. 3 suggests that $`k_{}k_{}`$ and, hence, that anisotropy is scale-independent. No theory mentioned in ยง1 agrees with our result. However, it is not clear from the figure whether or not the true anisotropy is a function of scale: although the contours show a linear relationship between $`k_{}`$ and $`k_{}`$, the relationship doesnโt mean that all individual eddies have the same major axis to minor axis ratio. As explained in next subsection, the $`rotation`$ of eddies by large-scale waves in the magnetic fields can distort the actual scaling relation and lead to the linear relationship shown in the figure.
Fig. 5 shows $`t_{phase}`$ as a function of wavenumber $`k`$. This time scale is defined as the average correlation time $`\mathrm{\Delta }t`$ such that Fourier components at $`๐ค=k`$ have a phase shift of $`60^o`$ with respect to the original phase. We plot the result for zero frequency modes (i.e. $`k_{}`$=0 modes). We see that $`t_{phase}k^1`$ (that is, $`k_{}^1`$). How can we interpret this result? If a turbulent structure, characterized by a wavenumber $`k`$, moves a distance $`l`$ then the phase is shifted by roughly $`kl`$, even if the eddy is unaffected by the motion. This implies that a large scale velocity $`V`$ will change the phase at a rate $`kV`$, so that $`t_{phase}k^1`$. This implies that our calculation of $`t_{phase}`$ is dominated by large scale motions rather than by the local cascade of energy, as long as the nonlinear cascade rate is proportional to $`k`$ to some power less than one, which is generally the case. This is an example of how large-scale fluctuations can complicate attempts to find physically meaningful scaling relations.
### 3.2 Rotation Effect
In the previous subsection, we showed that anisotropy appears to be scale-independent. In this subsection, we will show that this apparent scaling relation is an artifact caused by the Fourier transformation. That is, we will demonstrate that large-scale modes in the magnetic field can distort the actual scaling law at smaller scales when we perform a Fourier transformation. In this manner, a straightforward evaluation of anisotropy in Fourier space is strongly contaminated by the curvature of the large-scale magnetic fields and does not reflect the actual local anisotropy when anisotropy is more pronounced at smaller scales. Consequently, figures 3 and 4 are compatible with any scaling law that predicts that smaller eddies are more elongated.
Fig. 3 implies that eddies have anisotropic shapes: on average, eddies are stretched along the direction of $`๐_0`$. However, not all eddies are aligned along the large scale field. The elongation of an eddy is determined by its interaction with the $`local`$ magnetic field, not the background field. Since the large-scale magnetic field lines wander with respect to $`๐_0`$, all smaller scale eddies have similar angular distributions around $`๐_0`$. We illustrate this effect in Fig. 6.
In this way, we can explain the results of Goldreich and Sridhar (1995) and Matthaeus et al. (1998) simultaneously. Suppose that eddies are oriented according to the local field lines (Fig. 6a). Goldreich and Sridharโs result implies smaller eddies are relatively more elongated. When we perform a Fourier transformation and measure the ratio of $`k_{}/k_{}`$, what we actually measure is not the ratio of the minor axis to the major axis of individual eddies. Instead, because the direction of the local magnetic field varies according to location and Fourier transformation is none other than a (weighted) averaging process, we actually measure the ratio averaged over all possible orientation of the eddies (Fig. 6b). The Fourier transformation sees that $`L_1`$ (and $`l_1`$) is the minor axis and $`L_2`$ (and $`l_2`$) is the major axis of an eddy. The ratio of $`L_1/L_2`$ ($`=l_1/l_2`$) is determined by the degree of the wandering of the large scale magnetic field lines with respect to $`๐_0`$ and, therefore, the ratio is nearly constant for all eddies, regardless of their sizes and shapes. In fact, the ratio will depend on the tangent of the angle between $`๐_0`$ and $`๐`$. Since $`k_{}1/L_1`$ and $`k_{}1/L_2`$, the measured value of the ratio $`k_{}/k_{}`$ is nearly scale-independent. We expect the angle $`\theta `$ ($`90^o\theta _w`$) between $`๐_0`$ and $`๐`$ to be
$$\mathrm{tan}\theta =b/B_0.$$
(13)
Therefore, we have
$$\mathrm{sin}\theta =b/B=\mathrm{cos}\theta _w,$$
(14)
which is exactly the scaling relation found by Matthaeus et al.
## 4 Measuring Eddy Shapes
In this section, we analyze the shape of eddies in real space. As explained in the previous section, we assume that elongated eddies are aligned in the direction of the local magnetic fields. If we want to visualize the shape of eddies, we need to first identify the direction of the local magnetic fields. It is important to note that, although we use the term โlocal magnetic fieldsโ for simplicity, the fields are different from $`๐(๐ซ)`$. Let us consider an eddy of size $`l`$. The โlocal magnetic fieldโ of the eddy must act as the โlarge-scale magnetic fieldโ for the eddy. Therefore, the โlocal magnetic fieldsโ for eddies of size $`l`$, must be smoothly varying vector fields whose characteristic length of variation is $`>l`$. Note also that another eddy of size $`l^{}`$ ($`l`$) at the same location can have a slightly different direction for the โlocal magnetic field.โ The โlocal magnetic fieldsโ are functions of position ($`๐ซ`$) and eddy size ($`l`$). Then, how can we define the direction of the local magnetic fields? We implement 2 independent numerical algorithms to calculate the direction of the $`local`$ magnetic fields.
In the first method, the local magnetic fields are calculated by
$$๐_l=(๐(๐ซ_1)+๐(๐ซ_2))/2$$
(15)
and the second order structure functions for $`v`$ and $`b`$ are given by
$`F_2^v(R,z)=<|๐(๐ซ_1)๐(๐ซ_2)|^2>,`$ (16)
$`F_2^b(R,z)=<|๐(๐ซ_1)๐(๐ซ_2)|^2>,`$ (17)
where $`R=|\widehat{๐ณ}\times (๐ซ_2๐ซ_1)|,`$ $`z=\widehat{๐ณ}(๐ซ_2๐ซ_1)`$ and $`\widehat{๐ณ}=๐_l/|๐_l|.`$ That is, $`R`$ and $`z`$ are coordinates in a cylindrical coordinate system in which the z-axis is parallel to $`๐_l`$ (Fig. 7). $`๐(๐ซ)`$ and $`๐(๐ซ)`$ are the total and perturbed magnetic fields at a point $`๐ซ`$. As usual, brackets denotes a spatial average.
In Fig. 8 we plot the second order structure functions in z-R plane. The horizontal axis (z-axis) is parallel to $`๐_l`$. The contours reflect the shapes of the eddies. Eddies are clearly elongated along the local field lines, and smaller eddies are more elongated. The velocity and magnetic fields show different structure at large scales. However, their small scale structure is quite similar.
In Fig. 9, we plot R-intercepts and z-intercepts of the contours. The R-intercept and z-intercept of a given contour can be regarded as a measure of $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$ for the corresponding eddy scale<sup>4</sup><sup>4</sup>4We interpret $`\stackrel{~}{k}_{}`$ as the inverse of the major axis of eddies and $`\stackrel{~}{k}_{}`$ as that of the minor axis. Since the major axis is assumed to be parallel to the local magnetic fields, $`\stackrel{~}{k}_{}`$ is the parallel wavenumber with respect to the local magnetic field direction. On the other hand, in ยง3.1, $`k_{}`$ is parallel to $`๐_0`$. Therefore, $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$ in this section have different meanings from $`k_{}`$ and $`k_{}`$ in ยง3.1.. Fitting the results for velocity fields gives
$$v:\stackrel{~}{k}_{}\{\begin{array}{cc}\stackrel{~}{k}_{}^{0.69},\hfill & (256HB_00.5)\hfill \\ \stackrel{~}{k}_{}^{0.70},\hfill & (256HB_01)\hfill \\ \stackrel{~}{k}_{}^{0.73},\hfill & (256PB_01)\hfill \end{array}$$
(18)
On the other hand, for the magnetic fields we find
$$b:\stackrel{~}{k}_{}\{\begin{array}{cc}\stackrel{~}{k}_{}^{0.64},\hfill & (256HB_00.5)\hfill \\ \stackrel{~}{k}_{}^{0.50},\hfill & (256HB_01)\hfill \\ \stackrel{~}{k}_{}^{0.53},\hfill & (256PB_01)\hfill \end{array}$$
(19)
The velocity fields show good agreement with the scaling relation, $`\stackrel{~}{k}_{}\stackrel{~}{k}_{}^{2/3}`$, proposed by Goldreich and Sridhar (1995). The power indices are insensitive to the strength of $`B_0`$ or the form of viscosity (and diffusivity). However, the magnetic field shows different scaling behavior. When the external field is moderately strong (256H-$`B_00.5`$), the power index is very close to 2/3. On the other hand, for stronger external fields (256H-$`B_01`$ and 256P-$`B_01`$), the power indices are smaller than 2/3. The results are insensitive to the choice of viscosity (and diffusivity). It is not clear whether the existence of a separate scaling law for the magnetic field represents a physical effect not included in Goldreich and Sridharโs model, e.g. the first signs of small scale intermittency in the magnetic field distribution, or merely the failure of the numerical models used here to fully resolve the inertial range of strong MHD turbulence.
In the second method, we employ a completely different approach. We obtain the local large scale magnetic fields by filtering out the small scale magnetic fields:
$$๐_\sigma (๐ซ)=\underset{๐ซ^{}}{}๐(๐ซ^{})\varphi (|๐ซ๐ซ^{}|),$$
(20)
where $`\varphi (r)\mathrm{exp}(r^2/\sigma _r^2)`$ is a gaussian function. To determine the shape of small scale eddies (i.e. eddies smaller than the filter size, $`\sigma _r`$), we consider the following quantities:
$`๐ฏ_\sigma (๐ซ)=๐(๐ซ)๐_\sigma (๐ซ),`$ (21)
$`๐_\sigma (๐ซ)=๐(๐ซ)๐_\sigma (๐ซ),`$ (22)
where $`๐_\sigma `$ and $`๐_\sigma `$ are filtered fields (cf. eq. (20)). The fields $`๐ฏ_\sigma (๐ซ)`$ and $`๐_\sigma (๐ซ)`$ represent small scale fluctuation of velocity and magnetic fields, the shape of which can be regarded as an adequate approximation of small scale eddies. From these two fields we calculate the structure functions
$`F^v(R,z)=|๐ฏ_\sigma (๐ซ_2)๐ฏ_\sigma (๐ซ_1)|,`$ (23)
$`F^b(R,z)=|๐_\sigma (๐ซ_2)๐_\sigma (๐ซ_1)|,`$ (24)
where $`R`$ and $`z`$ are similarly defined as in the first method with $`\widehat{๐ณ}๐_\sigma (๐ซ_1)`$ (Fig. 10).
In Fig. 11, we plot the results of the second method. We can clearly observe the flattening effect: When the filter size is large, eddies are less anisotropic. When filter size is small, eddies show highly anisotropic structure. In the figure, we plot only the results for magnetic fields. Velocity fields show similar trends. The thick contours represent $`F^b(R,z)=<|๐_\sigma |^2>^{1/2}`$. The values next to the thick contours are $`<|๐_\sigma |^2>^{1/2}`$.
This second method is useful for visualizing small scale eddies, but may not be as useful for quantitative analysis. The difficulty comes from the fact that there is no well defined eddy scale associated with filtered fields $`๐ฏ_\sigma (๐ซ)`$ and $`๐_\sigma (๐ซ)`$.
In this paper, we will not pursue quantitative analysis using the method 2. Instead, we just wish to point out that both methods describe the same scaling law: smaller eddies are relatively more stretched along the local magnetic field lines. In particular, the results from the first method are consistent with the relation $`\stackrel{~}{k}_{}\stackrel{~}{k}_{}^{2/3}`$ proposed by Goldreich and Sridhar (1995).
## 5 Discussion and Conclusions
Here we rederive the scaling law $`\stackrel{~}{k}_{}\stackrel{~}{k}_{}^{2/3}`$ in the framework of 3-wave interactions. Except for the use of the uncertainty principle, the work in this section is independent of Goldreich and Sridharโs derivation. As noted by Goldreich and Sridhar (1995), 3-wave interactions are an adequate proxy for wave-wave interactions of all orders in a strong MHD turbulence. As long as we assume the locality of interactions, it is pointless to distinguish $`k_{}`$ and $`k_{}`$ from $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$. Hence, for simplicity, we use $`k_{}`$ and $`k_{}`$ instead of $`\stackrel{~}{k}_{}`$ and $`\stackrel{~}{k}_{}`$ during the derivation.
Suppose the 3-dimensional energy spectrum is given by
$$E_3(k_{},k_{})|\widehat{๐}(๐ค)|^2k_{}^{2\alpha }f,$$
(25)
where $`\widehat{๐}(๐ค)`$ is the amplitude of the mode whose wavevector is $`๐ค`$ and $`f(u)`$ is a positive, symmetric function of $`u`$ (cf. equation (7) of Goldreich and Sridhar (1995)) which describes the power distribution as a function of eddy shape. If the $`width`$ (or, $`thickness`$) of the energy spectrum in the direction of $`k_{}`$ is $`k_{}^\beta `$, then we can write
$$E_3(k_{},k_{})k_{}^{2\alpha }f(k_{}/k_{}^\beta ).$$
(26)
If the $`width`$ is caused by the uncertainty principle ($`\mathrm{\Delta }t\mathrm{\Delta }\omega 1`$ with $`\mathrm{\Delta }tt_{cas}(l)`$ and $`\mathrm{\Delta }\omega k_{}`$), then
$$t_{cas}(l)k_{}^\beta .$$
(27)
Suppose the energy cascade rate $`ฯตv_l^2/t_{cas}(l)`$, where $`l=2\pi /k_{}`$, is scale-independent. Because $`v_l^2k_{}^{2\alpha }k_{}^2k_{}^\beta `$ ($`k_{}E(k_{})`$, $`E(k_{})=`$ 1-dimensional spectrum) and $`t_{cas}(l)k_{}^\beta `$, we have
$$k_{}^{2\alpha +2+2\beta }=const.$$
(28)
Therefore,
$$1\alpha +\beta =0.$$
(29)
Now, letโs pick up a mode at a wavevector $`๐ฉ`$ and consider nonlinear interactions with other wave modes. First, the strength of the interaction with another mode at $`๐ช`$ is $`p|\widehat{๐}(๐ฉ)||\widehat{๐}(๐ช)|`$. (This comes from the $`(\times ๐)\times ๐`$ term in equation (9)) Hereafter we assume $`p|๐ฉ|p_{}`$. Second, the number of interactions is $`p^2p^\beta `$. This is the number of modes $`near`$ $`๐ฉ`$. Here we use locality of 3-wave interactions. If the interactions are random, the net change of amplitude per unit time will be the strength of the interaction times the square root of the number of interactions, or
$$|\mathrm{\Delta }\widehat{๐}(๐ฉ)|p|\widehat{๐}(๐ฉ)||\widehat{๐}(๐ช)|(p^2p^\beta )^{1/2}.$$
(30)
Therefore, we have
$$t_{cas}|\widehat{๐}(๐ฉ)|/|\mathrm{\Delta }\widehat{๐}(๐ฉ)|p^2p^\alpha p^{\beta /2},$$
(31)
where we assumed $`pq`$. Equating this with $`t_{cas}p^\beta `$, we can write
$$\alpha 2=\beta /2.$$
(32)
From equations (29) and (32), we have
$$\alpha =5/3,\beta =2/3,$$
(33)
which is just the result of Goldreich and Sridhar (1995):
$$k_{}k_{}^{2/3}.$$
(34)
As a consequence, the 3-D energy spectrum becomes
$$E_3(k_{},k_{})k_{}^{10/3}f(k_{}/k_{}^{2/3})$$
(35)
and the corresponding 1-D spectrum is given by
$$E(k)k^{5/3}.$$
(36)
In summary, we have shown that the anisotropy of Alfvรฉnic turbulence depends on the spatial scales of eddies. In particular, our results confirm the claim by Goldreich and Sridhar (1995, 1997) that smaller eddies are relatively more elongated along the direction of the local magnetic field lines than larger ones. Quantitative measurements of the anisotropy using the velocity fields show good agreement with their proposed scaling law, $`k_{}k_{}^{2/3}`$ as long we interpret these wavenumbers as referring to the local magnetic field direction. However, when the external magnetic field is very strong, magnetic fields scale somewhat differently, showing a slightly more rapid increase in anisotropy at smaller scales.
It is important to note that the correct scaling laws depend on comparing the eddy shape to the local magnetic field direction.
As a final note, we wish to stress that our results are not in agreement with the IK theory. The IK theory is based on the assumption of isotropy in wavenumber space, which may be true when the external magnetic field is very weak or zero. However, even in these cases, the turbulence is globally isotropic, but locally very anisotropic. In this paper, we showed that eddies do show anisotropy and that the anisotropy is scale-dependent when there is a strong large scale field (which should apply to very small scales within any MHD turbulence cascade). On the other hand, our results are consistent with Goldreich and Sridharโs theory of strong MHD turbulence. More precisely, if we consider the ratio of hydrodynamic to Alfvรฉnic rates, that is $`(kv_k/\stackrel{~}{k}_{}V_A)`$, we find from equations (18) and (19) that
$$\frac{kv_k}{\stackrel{~}{k}_{}V_A}\stackrel{~}{k}_{}^{0.30.5}v_k,$$
(37)
where $`k\stackrel{~}{k}_{}`$. From Fig. 2 we see that for the inertial range this implies a ratio which is either constant or increasing with wavenumber. A constant ratio is predicted by Goldreich and Sridharโs model. The IK model predicts a slow decline.
This work was partially supported by National Computational Science Alliance under CTS980010N and utilized the NCSA SGI/CRAY Origin2000. |
warning/0003/nucl-th0003065.html | ar5iv | text | # Realistic Calculation of the โโข๐โข๐ Astrophysical Factor
## Abstract
The astrophysical factor for the proton weak capture on <sup>3</sup>He is calculated with correlated-hyperspherical-harmonics bound and continuum wave functions corresponding to a realistic Hamiltonian consisting of the Argonne $`v_{18}`$ two-nucleon and Urbana-IX three-nucleon interactions. The nuclear weak charge and current operators have vector and axial-vector components, that include one- and many-body terms. All possible multipole transitions connecting any of the $`p^3`$He S- and P-wave channels to the <sup>4</sup>He bound state are considered. The $`S`$-factor at a $`p^3`$He center-of-mass energy of $`10`$ keV, close to the Gamow-peak energy, is predicted to be $`10.1\times 10^{20}`$ keV b, a factor of five larger than the standard-solar-model value. The P-wave transitions are found to be important, contributing about 40 % of the calculated $`S`$-factor.
preprint: JLAB-THY-00-09
Recently, there has been a revival of interest in the reaction <sup>3</sup>He($`p`$,$`e^+\nu _e`$)<sup>4</sup>He . This interest has been spurred by the Super-Kamiokande collaboration measurements of the energy spectrum of electrons recoiling from scattering with solar neutrinos . At energies larger than 14 MeV more recoil electrons have been observed than expected on the basis of standard-solar-model (SSM) predictions . The $`hep`$ process, as the proton weak capture on <sup>3</sup>He is known, is the only source of solar neutrinos with energies larger than 15 MeVโtheir end-point energy is about 19 MeV. This fact has naturally led to questions about the reliability of the currently accepted SSM value for the astrophysical factor at zero energy, $`2.3\times 10^{20}`$ keV b . In particular, Bahcall and Krastev have shown that a large enhancement, by a factor in the range 20โ30, of the SSM $`S`$-factor value given above would essentially fit the observed excess of recoiling electrons.
The theoretical description of the $`hep`$ process, as well as that of the neutron and proton radiative captures on deuteron and <sup>3</sup>He, constitute a challenging problem from the standpoint of nuclear few-body theory. Its difficulty can be appreciated by comparing the measured values for the cross section of thermal neutron radiative capture on <sup>1</sup>H, <sup>2</sup>H, and <sup>3</sup>He. Their respective cross sections are: $`334.2\pm 0.5`$ mb , $`0.508\pm 0.015`$ mb , and $`0.055\pm 0.003`$ mb . Thus, in going from $`A`$=2 to 4 the cross section has dropped by almost four orders of magnitude. These processes are induced by magnetic-dipole transitions between the initial two-cluster state in relative S-wave and the final bound state. In fact, the inhibition of the $`A`$=3 and 4 captures has been understood for a long time : the <sup>3</sup>H and <sup>4</sup>He states are approximate eigenstates of the magnetic dipole operator $`๐`$, and consequently matrix elements of $`\mu _z`$ between $`nd`$ ($`n^3`$He) and <sup>3</sup>H (<sup>4</sup>He) vanish (approximately) due to orthogonality. This orthogonality argument fails in the case of the deuteron, since then $`\mu _z`$ can connect the large S-wave component of the deuteron to an isospin $`T`$=1 <sup>1</sup>S<sub>0</sub> $`np`$ state.
This quasi-orthogonality, while again invalid in the case of the proton weak capture on protons , is also responsible for inhibiting the $`hep`$ process. Both these reactions are induced by the Gamow-Teller operator, which differs from the (leading) isovector spin part of the magnetic dipole operator essentially by an isospin rotation. As a result, the $`hep`$ weak capture and $`nd`$, $`pd`$ and $`n^3`$He radiative captures are extremely sensitive to: i) the D-state admixtures generated by tensor interactions, and ii) many-body terms in the electroweak current operator. For example, many-body current contributions provide, respectively, 50 % and over 90 % of the calculated $`pd`$ and $`n^3`$He cross sections at very low energies.
In this respect, the $`hep`$ weak capture is a particularly delicate reaction, for two additional reasons: firstly and most importantly, the one- and many-body current contributions are comparable in magnitude, but of opposite sign ; secondly, many-body axial currents, specifically those arising from excitation of $`\mathrm{\Delta }`$ isobars which give the dominant contribution, are model dependent . This destructive interference between one- and many-body currents also occurs in the $`n^3`$He (โ$`hen`$โ) radiative capture , with the difference that there the leading components of the many-body currents are model independent, and give a much larger contribution than that associated with the one-body current.
The cancellation in the $`hep`$ process between the one- and two-body matrix elements has the effect of enhancing the importance of P-wave capture channels. Indeed, one of the results of the present work is that these channels give about 40 % of the $`S`$-factor calculated value. That the $`hep`$ process could proceed as easily through P- as S-wave capture was not not sufficiently appreciated in all earlier studies of this reaction we are aware of, with the exception of Ref. , in which Horowitz suggested, on the basis of a very simple one-body reaction model, that the <sup>3</sup>P<sub>0</sub> channel may be important.
Most of the earlier studies had attempted to relate the matrix element of the axial current occurring in the $`hep`$ capture to that of the electromagnetic current in the $`hen`$ capture, exploiting (approximate) isospin symmetry. This approach led, however, to $`S`$-factor values ranging from $`3.7`$ to $`57`$, in units of $`10^{20}`$ keV b. In an attempt to reduce the uncertainties in the predicted values for both the radiative and weak capture rates, ab initio microscopic calculations of these reactions were performed in the early nineties , using variational wave functions corresponding to a realistic Hamiltonian, and a nuclear electroweak current consisting of one- and many-body components. These studies showed that inferring the $`hep`$ $`S`$-factor from the measured $`hen`$ cross section can be misleading, because of different initial-state interactions in the $`n^3`$He and $`p^3`$He channels, and because of the large contributions associated with the two-body components of the electroweak current operator, and their destructive interference with the one-body current contributions.
The significant progress made in the last few years in the modeling of two- and three-nucleon interactions and the nuclear weak current, and the description of the bound and continuum four-nucleon wave functions, have prompted us to re-examine the $`hep`$ reaction. In the present work we briefly summarize the salient points in the calculation, and report our results for the $`S`$-factor in the energy range 0โ10 keV. An exhaustive account of this study , however, will be published elsewhere.
The cross section for the <sup>3</sup>He($`p`$,$`e^+\nu _e`$)<sup>4</sup>He reaction at a c.m. energy $`E`$ is written as
$`\sigma (E)={\displaystyle }`$ $`2\pi \delta \left(\mathrm{\Delta }m+E{\displaystyle \frac{q^2}{2m_4}}E_eE_\nu \right){\displaystyle \frac{1}{v_{\mathrm{rel}}}}`$ (2)
$`\times {\displaystyle \frac{1}{4}}{\displaystyle \underset{s_es_\nu }{}}{\displaystyle \underset{s_1s_3}{}}|f|H_W|i|^2{\displaystyle \frac{d๐ฉ_e}{(2\pi )^3}}{\displaystyle \frac{d๐ฉ_\nu }{(2\pi )^3}},`$
where $`\mathrm{\Delta }m=m+m_3m_4`$ = 19.8 MeV ($`m`$, $`m_3`$, and $`m_4`$ are the proton, <sup>3</sup>He, and <sup>4</sup>He rest masses, respectively), $`v_{\mathrm{rel}}`$ is the $`p^3`$He relative velocity, and the transition amplitude is given by
$$f|H_W|i=\frac{G_V}{\sqrt{2}}l^\sigma ๐ช;^4\mathrm{He}|j_\sigma ^{}(๐ช)|๐ฉ;p^3\mathrm{He}.$$
(3)
Here $`G_V`$ is the Fermi constant, $`๐ช=๐ฉ_e+๐ฉ_\nu `$, $`|๐ฉ;p^3\mathrm{He}`$ and $`|๐ช;^4\mathrm{He}`$ represent, respectively, the $`p^3`$He scattering state with relative momentum $`๐ฉ`$ and <sup>4</sup>He bound state recoiling with momentum $`๐ช`$, $`l_\sigma `$ is the leptonic weak current, $`l_\sigma =\overline{u}_\nu \gamma _\sigma (1\gamma _5)v_e`$ (the lepton spinors are normalized as $`v_e^{}v_e=u_\nu ^{}u_\nu =1`$), and $`j^\sigma (๐ช)`$ is the nuclear weak current, $`j^\sigma (๐ช)=(\rho (๐ช),๐ฃ(๐ช))`$. The dependence of the amplitude upon the spin projections of the leptons, proton and <sup>3</sup>He has been omitted for ease of presentation. Since the energies of interest are of the order of 10 keV or lessโthe Gamow-peak energy is 10.7 keV for the $`hep`$ reactionโit is convenient to expand the $`p^3`$He scattering state into partial waves, and perform a multipole decomposition of the nuclear weak charge, $`\rho (๐ช)`$, and current, $`๐ฃ(๐ช)`$, operators. Standard manipulations lead to
$$\frac{1}{4}\underset{s_es_\nu }{}\underset{s_1s_3}{}|f|H_W|i|^2=(2\pi )^2G_V^2L_{\sigma \tau }N^{\sigma \tau },$$
(4)
where the lepton tensor $`L^{\sigma \tau }`$ is written in terms of electron and neutrino four-velocities as $`L^{\sigma \tau }=\mathrm{v}_e^\sigma \mathrm{v}_\nu ^\tau +\mathrm{v}_\nu ^\sigma \mathrm{v}_e^\tau g^{\sigma \tau }\mathrm{v}_e\mathrm{v}_\nu +\mathrm{i}ฯต^{\sigma \alpha \tau \beta }\mathrm{v}_{e,\alpha }\mathrm{v}_{\nu ,\beta }`$, while the nuclear tensor is defined as
$$N^{\sigma \tau }\underset{s_1s_3}{}W^\sigma (๐ช;s_1s_3)W^\tau (๐ช;s_1s_3),$$
(5)
with
$`W^{\sigma =0,3}(๐ช;s_1s_3)`$ $`=`$ $`{\displaystyle \underset{LSJ}{}}X_0^{LSJ}(\widehat{๐ช};s_1s_3)T_J^{LSJ}(q),`$ (6)
$`W^{\sigma =\lambda }(๐ช;s_1s_3)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}{\displaystyle \underset{LSJ}{}}X_\lambda ^{LSJ}(\widehat{๐ช};s_1s_3)`$ (8)
$`\left[\lambda M_J^{LSJ}(q)+E_J^{LSJ}(q)\right],`$
where $`\lambda =\pm 1`$ denote spherical components. The functions $`X_{\lambda =0,\pm 1}`$ depend upon the direction $`\widehat{๐ช}`$, and the proton and <sup>3</sup>He spin projections $`s_1`$ and $`s_3`$ (note that the quantization axis for the hadronic states is taken along $`\widehat{๐ฉ}`$, the direction of the $`p^3`$He relative momentum), while $`T_J^{LSJ}`$=$`C_J^{LSJ}`$ or $`L_J^{LSJ}`$ for $`\sigma `$=0 or 3. The quantities $`C_J^{LSJ}`$, $`L_J^{LSJ}`$, $`E_J^{LSJ}`$ and $`M_J^{LSJ}`$ are the reduced matrix elements (RMEs) of the Coulomb $`(C_\mathrm{}\mathrm{}_z)`$, longitudinal $`(L_\mathrm{}\mathrm{}_z)`$, transverse electric $`(E_\mathrm{}\mathrm{}_z)`$, and transverse magnetic $`(M_\mathrm{}\mathrm{}_z)`$ multipole operators between the initial $`p^3`$He state with orbital angular momentum $`L`$, channel spin $`S`$ ($`S`$=0,1), and total angular momentum $`J`$, and final <sup>4</sup>He state. The present study includes S- and P-wave capture channels, i.e. the <sup>1</sup>S<sub>0</sub>, <sup>3</sup>S<sub>1</sub>, <sup>3</sup>P<sub>0</sub>, <sup>1</sup>P<sub>1</sub>, <sup>3</sup>P<sub>1</sub>, and <sup>3</sup>P<sub>2</sub> states, and retains all contributing multipoles connecting these states to the $`J^\pi `$=0<sup>+</sup> ground state of <sup>4</sup>He.
The bound- and scattering-state wave functions are obtained variationally with the correlated-hyperspherical-harmonics (CHH) method, developed in Refs. . The nuclear Hamiltonian consists of the Argonne $`v_{18}`$ two-nucleon and Urbana-IX three-nucleon interactions. This realistic Hamiltonian, denoted as AV18/UIX, reproduces the experimental binding energies and charge radii of the trinucleons and <sup>4</sup>He in โexactโGreenโs function Monte Carlo (GFMC) calculations . The binding energy of <sup>4</sup>He calculated with the CHH method is within 1 % of that obtained with the GFMC method. The accuracy of the CHH method to calculate scattering states has been successfully verified in the case of three-nucleon systems, by comparing results for a variety of $`Nd`$ scattering observables obtained by a number of groups using different techniques . Studies along similar lines to assess the accuracy of the CHH solutions for the four-nucleon continuum have already begun.
The CHH predictions for the $`n^3`$H total elastic cross section and coherent scattering length have been found to be in excellent agreement with the corresponding experimental values. The $`n^3`$H cross section is known over a rather wide energy range, and its extrapolation to zero energy is not problematic . The situation is different for the $`p^3`$He channel, for which the singlet and triplet scattering lengths $`a_\mathrm{s}`$ and $`a_\mathrm{t}`$ have been determined from effective range extrapolations of data taken above 1 MeV, and are therefore somewhat uncertain, $`a_\mathrm{s}=(10.8\pm 2.6)`$ fm and $`a_\mathrm{t}=(8.1\pm 0.5)`$ fm or $`(10.2\pm 1.5)`$ fm . Nevertheless, the CHH results are close to the experimental values above: the AV18/UIX Hamiltonian predicts $`a_\mathrm{s}=11.5`$ fm and $`a_\mathrm{t}=9.13`$ fm. At low energies (below 4 MeV) $`p^3`$He elastic scattering proceeds mostly through S- and P-wave channels, and the CHH predictions, based on the AV18/UIX model, for the differential cross section are in good agreement with the experimental data.
The nuclear weak current has vector and axial-vector parts, with corresponding one- and many-body components. The one-body components have the standard expressions obtained from a non-relativistic reduction of the covariant single-nucleon vector and axial-vector currents, including terms proportional to $`1/m^2`$. The two-body weak vector currents are constructed from the isovector two-body electromagnetic currents in accordance with the conserved-vector-current hypothesis, and consist of โmodel-independentโand โmodel-dependentโterms. The model-independent terms are obtained from the nucleon-nucleon interaction, and by construction satisfy current conservation with it. The leading two-body weak vector current is the โ$`\pi `$-likeโoperator, obtained from the isospin-dependent spin-spin and tensor nucleon-nucleon interactions. The latter also generate an isovector โ$`\rho `$-likeโcurrent, while additional isovector two-body currents arise from the isospin-independent and isospin-dependent central and momentum-dependent interactions. These currents are short-ranged, and numerically far less important than the $`\pi `$-like current. With the exception of the $`\rho `$-like current, they have been neglected in the present work. The model-dependent currents are purely transverse, and therefore cannot be directly linked to the underlying two-nucleon interaction. The present calculation includes the currents associated with excitation of $`\mathrm{\Delta }`$ isobars which, however, are found to give a rather small contribution in weak-vector transitions, as compared to that due to the $`\pi `$-like current. The $`\pi `$-like and $`\rho `$-like contributions to the weak vector charge operator have also been retained in the present study.
The leading many-body terms in the axial current due to $`\mathrm{\Delta }`$-isobar excitation are treated non-perturbatively in the transition-correlation-operator (TCO) scheme, originally developed in Ref. and further extended in Ref. . In the TCO schemeโessentially, a scaled-down approach to a full $`N`$$`+`$$`\mathrm{\Delta }`$ coupled-channel treatmentโthe $`\mathrm{\Delta }`$ degrees of freedom are explicitly included in the nuclear wave functions. The axial charge operator includes, in addition to $`\mathrm{\Delta }`$-excitation terms (which, however, are found to be unimportant ), the long-range pion-exchange term , required by low-energy theorems and the partially-conserved-axial-current relation, as well as the (expected) leading short-range terms constructed from the central and spin-orbit components of the nucleon-nucleon interaction, following a prescription due to Riska and collaborators .
The largest model dependence is in the weak axial current. To minimize it, the poorly known $`N`$$`\mathrm{\Delta }`$ transition axial coupling constant $`g_A^{}`$ has been adjusted to reproduce the experimental value of the Gamow-Teller matrix element in tritium $`\beta `$-decay . While this procedure is model dependent, its actual model dependence is in fact very weak, as has been shown in Ref. .
The calculation proceeds in two steps : first, the matrix elements of $`\rho (๐ช)`$ and $`๐ฃ(๐ช)`$ between the initial $`p^3`$He $`LSJJ_z`$ states and final <sup>4</sup>He are calculated with Monte Carlo integration techniques; second, the contributing RMEs are extracted from these matrix elements, and the cross section is calculated by performing the integrations over the electron and neutrino momenta in Eq. (2) numerically, using Gauss points.
The results for the $`S`$-factor, defined as $`S(E)=E\sigma (E)\mathrm{exp}(4\pi \alpha /v_{\mathrm{rel}})`$ ($`\alpha `$ is the fine structure constant), at $`p^3`$He c.m. energies of $`0`$, $`5`$, and $`10`$ keV are reported in Table I. In the table, the column labelled S includes both the <sup>1</sup>S<sub>0</sub> and <sup>3</sup>S<sub>1</sub> channel contributions, although the former are at the level of a few parts in $`10^3`$. The energy dependence is rather weak: the value at $`10`$ keV is only about 4 % larger than that at $`0`$ keV. The P-wave capture states are found to be important, contributing about 40 % of the calculated $`S`$-factor. However, the contributions from D-wave channels are expected to be very small. We have verified explicitly that they are indeed small in <sup>3</sup>D<sub>1</sub> capture. The many-body axial currents associated with $`\mathrm{\Delta }`$ excitation play a crucial role in the (dominant) <sup>3</sup>S<sub>1</sub> capture, where they reduce the $`S`$-factor by more than a factor of four. Thus the destructive interference between the one- and many-body current contributions, first obtained in Ref. , is confirmed in the present study. The (suppressed) one-body contribution comes mostly from transitions involving the D-state components of the <sup>3</sup>He and <sup>4</sup>He wave functions, while the many-body contributions are predominantly due to transitions connecting the S-state in <sup>3</sup>He to the D-state in <sup>4</sup>He, or viceversa.
It is important to stress the differences between the present and all previous studies. Apart from ignoring, or at least underestimating, the contribution due to P-waves, the latter only considered the long-wavelength form of the weak multipole operators, namely, their $`q`$=$`0`$ limit. In <sup>3</sup>P<sub>0</sub> capture, for example, only the $`C_0`$-multipole, associated with the weak axial charge, survives in this limit, and the corresponding $`S`$-factor is calculated to be $`2.2\times 10^{20}`$ keV b, including two-body contributions. However, when the transition induced by the longitudinal component of the axial current (via the $`L_0`$-multipole, which vanishes at $`q`$=$`0`$) is also taken into account, the $`S`$-factor becomes $`0.82\times 10^{20}`$ keV b, because of a destructive interference between the $`C_0`$ and $`L_0`$ RMEs. Thus use of the long-wavelength approximation in the calculation of the $`hep`$ $`S`$-factor leads to inaccurate results.
Finally, besides the differences listed above, the present calculation also improves that of Ref. in a number of other important respects: firstly, it uses accurate CHH wave functions, corresponding to the last generation of realistic interactions; secondly, the model for the nuclear weak current has been extended to include the axial charge as well as the vector charge and current operators. Thirdly, the one-body operators now take into account the $`1/m^2`$ relativistic corrections, which were previously neglected. In <sup>3</sup>S<sub>1</sub> capture, for example, these terms increase by 25 % the dominant (but suppressed) $`L_1`$ and $`E_1`$ RMEs calculated with the (lowest order) Gamow-Teller operator. These improvements in the treatment of the one-body axial current indirectly affect also the contributions of the $`\mathrm{\Delta }`$-excitation currents , because of the procedure used to determine the coupling constant $`g_A^{}`$.
To conclude, we have carried out a realistic calculation of the $`hep`$ reaction, predicting a value for the $`S`$-factor five times larger than that used in the SSM. This enhancement, while very significant, is far smaller than that required by fits to the Super-Kamiokande data. Although the present result is inherently model dependent, it is unlikely the model dependence be so large to accomodate a drastic increase in the prediction obtained here.
The authors wish to thank J.F. Beacom, J. Carlson, V.R. Pandharipande, D.O. Riska, P. Vogel, and R.B. Wiringa for useful discussions. The support of the U.S. Department of Energy under contract number DE-AC05-84ER40150 is gratefully acknowledged by L.E.M. and R.S. |
warning/0003/nlin0003021.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Stable quasiperiodic (QP) planforms have been first discovered as equilibrium patterns in metallic alloys . The simplest and most important types of the QP planforms are the ten-, eight-, and twelvefold ones (note an argument according to which only these types of the two-dimensional quasicrystals may occur in physical systems ). A $`2n`$-fold pattern is a superposition of $`n4`$ spatial harmonics based on a set of equal-length wave vectors $`๐ค_j`$ with the equal angles $`\pi /n`$ between $`๐ค_j`$ and $`๐ค_{j+1}`$ (more general aperiodic patterns with unequal wave vector lengths and/or unequal angles between the vectors may also exist, but they have probably never been considered, except for the case of unequilateral hexagons with $`n=3`$ ).
Then, it was predicted that the $`2n`$-fold structures of different particular types may exist as dynamical planforms of a stationary form in nonequilibrium systems, such as thermal convection and the like . Later, the predicted pattern (a twelvefold one, in particular) was indeed experimentally observed, first, as a stable dynamical nonequilibrium planform in the Faraday ripples (a large-aspect-ratio liquid layer subject to high-frequency shaking in the vertical direction), and recently in an optical cell filled with Na vapors . Nevertheless, experimental generation of dynamical QP patterns is far from being straightforward; in particular, in the work it was necessary to shake the liquid layer with a two-frequency quasiperiodic force at a specially selected ratio between the two frequencies. These experimental difficulties are related to the general fact that, in terms of the corresponding coupled Ginzburg-Landau (GL) equations for amplitudes of the spatial modes, a superposition of which gives rise to the pattern, a formal solution for the QP planforms always exists, but it may be stable only in a relatively narrow parametric region .
An objective of this work is to put forward a much easier possibility of generating eight- and twelvefold QP planforms of a finite but large size (a stripe) in โnormalโ systems, where direct generation of the quasicrystallic patterns does not seem feasible. To this end, we notice that the set of the wave vectors on which the eight- or twelvefold QP is based may be regarded, in an obvious way, as a superposition of two half-sets of $`2`$ or $`3`$ vectors, see Fig. 1 below. Each half-set, in turn, may give rise to a usual periodic pattern consisting of square cells (SC) or hexagonal cells, respectively. Thus, a natural way to generate a stripe filled with the QP pattern in a generic system (e.g., a convective layer), in which a stable QP pattern is not available, but stable periodic SC and/or hexagonal patterns do exist, is to produce it as a transient layer (โdomain wallโ ) between two large domains filled with the periodic cells, the (half-) sets of the two or three wave vectors in the two domains being oriented under the angle, respectively, $`45`$ or $`30`$ degrees relative to each other. The pattern in the transient layer will then be a superposition of the two periodic patterns, having a full set of the four or six wave vectors, respectively. The so generated stripe will feature a QP pattern, provided that its width is essentially larger than the wavelength (a size of the elementary cell in the corresponding periodic pattern). It is known that the latter condition can be achieved, under special but not unrealistic conditions, for domain walls in the nonequilibrium systems .
In fact, it may be simpler to generate, following this way, a twelvefold QP pattern between two hexagonal domains, as it is usually much easier to find a stable hexagonal structure than a stable SC one . However, the theoretical analysis of the transient layer between two SC domains is much simpler, therefore in this work we concentrate on the latter case. In any case, it should be stressed that, although stable square-cell patterns are rarer than their counterparts in the form of hexagons or rolls, examples of stable square cellular planforms are known in gas flames , thermal convection , and in some optical systems . The interest to the SC planforms has been recently revived by the discovery of this pattern in a double-layer Marangoni convection (see the e-print and references therein).
The rest of the paper is organized as follows. In section 2, we formulate the model. Some analytical results, which predict, in particular, that the transient layer between the two SC patterns has a complex structure, consisting of three layers, two broad ones and a narrow sublayer sandwiched between them, are obtained in section 3. In section 4, we display results of a direct numerical solution of the stationary real GL equations, which comply with the analytical predictions. The paper is concluded by section 5.
## 2 The Model
In this work, we assume a simplest configuration, shown in Fig. 1, that is going to give rise to the QP transient layer. The layer is parallel to the $`y`$ axis, and the (half) sets of the two wave vectors are chosen so that in the left domain the vectors $`๐ค_{1,2}`$ are parallel to the $`x`$ and $`y`$ axes, while in the right domain both vectors $`๐ค_{3,4}`$ have the angle $`45`$ degrees relative to the axes. Accordingly, the complex wave field describing the spatial distribution of physical variables is assumed to be
$$u(๐ซ,t)=\underset{j=1}{\overset{4}{}}B_j(๐ซ,t)\mathrm{exp}\left(i๐ค_j๐ซ\right),$$
(1)
where $`๐ซ`$ is the two-dimensional coordinate, and $`B_j(๐ซ,t)`$ are slowly varying amplitude functions.
In an experiments, the chosen configuration can be created, at least in a part of the system, by means of specially selected boundary conditions, which, in turn, are imposed by the sidewalls of the experimental cell. Although we do not analyze the sidewall boundary conditions in this work, it is obvious that, to support the configuration that we consider, one will need to have a large-aspect-ratio cell with two sidewalls forming an angle $`45`$ degrees, which is quite possible. More general configurations, with different orientations of the wave vectors relative to the layer between the two domains, can be considered similarly to what is done below, but their technical treatment will be more cumbersome.
A usual approach to the description of spatially nonuniform patterns of the domain-wall type is based on a system of coupled real Ginzburg-Landau (GL) equations for the slowly varying amplitudes, assuming that they do not depend on the coordinate $`y`$ running along the stripe (wall), which is a reasonable approximation if the stripe is long enough . Then, as it was shown in the mentioned works, the effective diffusion coefficient (the one in front of the term $`(B_j)_{xx}`$) in each GL equation is
$$D_j=\left(k_j^{(x)}\right)^2,$$
(2)
where $`k_j^{(x)}`$ is the $`x`$-component of the vector. For the configuration shown in Fig. 1, the system of the real GL equations can be easily cast into a final form
$`A_t`$ $`=`$ $`(1/2)A_{xx}+AA^32\left(g_2A^2+g_1B_1^2+g_1B_2^2\right)A,`$ (3)
$`\left(B_1\right)_t`$ $`=`$ $`\left(B_1\right)_{xx}+B_1B_1^32\left(2g_1A^2+g_2B_2^2\right)B_1,`$ (4)
$`\left(B_2\right)_t`$ $`=`$ $`B_2B_2^32\left(2g_1A^2+g_2B_1^2\right)B_2,`$ (5)
where $`B_{1,2}`$ are the amplitudes corresponding to the wave vectors $`๐ค_{1,2}`$ in the left domain (Fig. 1), and, using the obvious symmetry of the configuration shown in Fig. 1, we have set $`B_3=B_4A`$.
Note that the diffusion coefficient in Eq. (5) is zero due to Eq. (2). Generally speaking, in this case one should take into account a higher-order derivative term $`\left(B_2\right)_{xxxx}`$ . However, this is not necessary while the diffusion terms do not vanish in the two other equations (see details below).
In Eqs. (3) through (5), the linear gain coefficient and the coefficient of the nonlinear self-interaction of the spatial mode are normalized to be $`1`$, $`g_1`$ and $`g_2`$ being the coefficients of the nonlinear interaction between the modes with the angles $`\alpha =45`$ and $`90`$ degrees between their carrier wave vectors. Normally, the nonlinear interaction coefficient decreases with the increase of $`\alpha `$, so that
$$g_2<g_1<1.$$
(6)
The necessary and sufficient stability conditions for the SC pattern are well known ,
$$g_2<\frac{1}{2},g_1>\frac{1}{4}\left(1+2g_2\right),$$
(7)
while the conditions providing for stability of the eightfold QP planform are
$$g_2<\frac{1}{2},g_1<\frac{1}{4}\left(1+2g_2\right).$$
(8)
An obvious feature of the two sets of the stability conditions is their incompatibility, i.e., SC and QP can never be stable simultaneously.
In fact, the inequality $`g_2<1/2`$ in the set of the conditions (7) is a cause for the relative rarity of stable SC planforms, as, despite the general property (6), the actual dependence $`g(\alpha )`$ is usually weak, so that $`g_2`$ is not essentially smaller than $`1`$. Nevertheless, the full set of the SC stability conditions (7) can be satisfied in the above-mentioned physical systems.
A consequence of the conditions (8) necessary for the stability of the QP pattern is $`g_1<1/2`$, which, with regard to Eq. (6), makes the stability of the QP pattern still less feasible than that of the SC one. An objective of this work is to propose a way to produce a transient QP pattern between two stable SC domains with different orientations (Fig. 1) in the case when the uniform QP planform is unstable. Note that a similar approach is known as a way to generate of a broad stripe of a SC pattern between two domains of orthogonally oriented rolls in the case when the uniform SC pattern is unstable, while the rolls are stable .
In the case when the SC stability conditions (7) are met, it is quite reasonable to assume that the coefficients $`g_{1,2}`$ are close to $`1/2`$ (because it is physically implausible to have $`g_{1,2}`$ much smaller than $`1/2`$), which suggests to present them as
$$g_{1,2}\frac{1}{2}\left(1\mu _{1,2}\right),|\mu _{1,2}|1.$$
(9)
As it will be seen below, the smallness of $`\mu _{1,2}`$ naturally provides for the strip of the QP pattern to be broad, which is exactly the condition justifying the consideration of the transient-layer patterns. Note that $`\mu _2`$ must be positive according to Eq. (7), while $`\mu _1`$ may formally have either sign. However, it will be shown below that the necessary solution does not exist if $`\mu _1<0`$.
Below, we will also use a parameter
$$m2\mu _1/\mu _2,$$
(10)
which is, generally, $`1`$. In terms of $`m`$, the second stability condition (7) takes a very simple form, $`m<1`$.
Eqs. (3), (4), and (5) must be supplemented by boundary conditions (b.c.) to guarantee that, at $`x\pm \mathrm{}`$ (recall the system is formally assumed to be infinitely large), the pattern considered asymptotically coincides with either of the two SC planforms composed of the modes $`B_1`$ and $`B_2`$ or $`B_3=B_4A`$. Obviously, this implies
$`\underset{x+\mathrm{}}{lim}A(x)`$ $`=`$ $`1/\sqrt{2\mu _2}A_{lim},\underset{x+\mathrm{}}{lim}B_{1,2}(x)=0,`$ (11)
$`\underset{x\mathrm{}}{lim}A(x)`$ $`=`$ $`0,\underset{x\mathrm{}}{lim}B_{1,2}(x)=A_{lim}.`$ (12)
Formally, this set may seem overdetermined, as we add six b.c. to the fourth-order system of Eqs. (3) and (4) (Eq. (5) contains no $`x`$-derivatives). However, it is easy to check that the seemingly superfluous b.c. for $`B_2`$ are nothing else but direct corollaries of the four legitimate b.c. for $`A`$ and $`B_1`$.
## 3 Analytical Results
A stationary version ($`/t=0`$) of Eqs. (3), (4) and (5) can be essentially simplified, as in this case Eq. (5) becomes just an algebraic relation, that has two solutions: $`B_2=0`$, or
$$B_2^2=1\left(2m\mu _2\right)A^2\left(1\mu _2\right)B_1^2,$$
(13)
$`m`$ being the parameter defined by Eq. (10). First, we consider the case when the expression (13) holds (obviously, it may hold as long as it yields $`B_2^2>0`$). Substituting it into the stationary versions of Eqs. (3) and (4), we obtain
$`\mu _2^1A^{\prime \prime }+mA\left[\left(2\left(2m1\right)m^2\mu _2\right)A^2+\left(2m\mu _2\right)B_1^2\right]A`$ $`=`$ $`0,`$ (14)
$`\mu _2^1B_1^{\prime \prime }+B_1\left[\left(2\mu _2\right)B_1^2+\left(2m\mu _2\right)A^2\right]B_1`$ $`=`$ $`0,`$ (15)
the prime standing for $`d/dx`$. If, instead of Eq. (13), we take $`B_2=0`$, the stationary equations take the form
$`A^{\prime \prime }+2A\left[2\left(2\mu _2\right)A^2+\left(2m\mu _2\right)B_1^2\right]A`$ $`=`$ $`0,`$ (16)
$`B_1^{\prime \prime }+B_1\left[B_1^2+\left(2m\mu _2\right)A^2\right]B_1`$ $`=`$ $`0.`$ (17)
It is noteworthy that, although the SC patterns may be stable at $`\mu _1<0`$, i.e., $`m<0`$, Eq. (14) cannot have a solution for $`A(x)`$ exponentially decaying at $`x\mathrm{}`$ (see the b.c. (12)) unless $`m>0`$, hence the present problem has no solution with $`m<0`$.
Obviously, a solution to Eqs. (14) and (15) can satisfy, with regard to Eq. (13), the b.c. (12). However, the same set of equations (14) and (15) cannot satisfy the b.c. (11): setting $`B_1=0`$ and $`A=\mathrm{const}`$, one obtains from Eq. (14)
$$A^2=A_0^2\frac{m}{2\left(2m1\right)m^2\mu _2},$$
(18)
which is obviously different from the necessary limit value $`A_{lim}^21/\left(2\mu _2\right)`$. Moreover, the value of $`B_2^2`$ corresponding to $`A_0^2`$ as per Eq. (13) is different from zero, which also violates the b.c. (11).
In fact, the stationary state corresponding to the asymptotic value (18) with $`B_20`$ is another uniform pattern, which is a superposition of three spatial harmonics. This pattern is periodic in one spatial direction and quasiperiodic in the other one. As it was demonstrated in , this pattern is always dynamically unstable, hence a solution having it as an asymptotic state is physically irrelevant.
However, a solution to Eqs. (14) and (15) makes sense as long as it provides for $`B_2^2>0`$ according to Eq. (13). Taking the asymptotic state $`A^2=A_0^2`$, $`B_1^2=0`$, one finds that it gives rise to negative $`B_2^2`$ exactly in the case $`m<1`$, which is considered in this work, as this is the case when the SC pattern is stable, see above. Thus, Eqs. (14) and (15) should be used in the region $`\mathrm{}<x<x_0`$, where, by definition, $`x=x_0`$ is a point at which $`B_2^2\left(x\right)`$, as given by Eq. (13), vanishes, i.e.,
$$\left(2m\mu _2\right)A^2(x_0)+\left(1\mu _2\right)B_1^2(x_0)=1.$$
(19)
At the point $`x=x_0`$, one must switch from Eqs. (14) and (15) to Eqs. (16) and (17), setting $`B_20`$ at $`xx_0`$. The continuity dictates to take the values $`A(xx_0^{})`$ and $`B_1(xx_0^{})`$ as the b.c. to Eqs. (16) and (17) at $`x=x_0`$, the b.c. at $`x=+\mathrm{}`$ being fixed by Eqs. (11).
The transition from Eqs. (14) and (15) to Eqs. (16) and (17) at $`x=x_0`$ provides for the continuity of all the functions $`A(x)`$ and $`B_{1,2}(x)`$, but their first derivatives suffer a jump at $`x=x_0`$. In fact, if the above-mentioned fourth-derivative term is added to Eq. (5), the jump of the derivative will be smoothed down in a narrow boundary layer. It should be stressed that the presence of the jump does not violate the applicability of the description in terms of the GL equations; the only problem is the absence of a detailed description of the narrow boundary layer in which the jump is smoothed by the fourth-order derivative term, but details of the inner structure of the narrow layer do not affect the global picture.
Thus, the transient layer between the two SC domains consists of two sublayers, described, respectively, by Eqs. (14) and (15), and by Eqs. (16) and (17), with the discontinuity of the derivative between them. The first (left) sublayer contains all the four spatial harmonics, hence is locally seems as a QP pattern. An important finding is that, in the case of small $`\mu _2`$ considered here, the width of this sublayer is large, scaling $`\mu _2^{1/2}`$, according to Eqs. (14) and (15). The possibility to produce a broad QP stripe justifies all the consideration of the transient layer. The second (right) sublayer contains only three spatial harmonics, as $`B_20`$ in it, hence it locally looks like a pattern quasiperiodic only in one direction ), and its width is $`1`$ (i.e., not specifically large), according to Eqs. (16) and (17).
This qualitative analysis of the transient layerโs structure will be corroborated and illustrated by direct numerical results displayed in the next section. However, before using the numerical methods, one can notice that the structure of the second (right) sublayer can be described in an approximate analytical form if we consider a special case when the value $`A(x=x_0)`$ at the internal boundary between the two sublayers is already close to the asymptotic value $`A_{lim}`$, and, accordingly, the value $`B_1^2`$ is small (in other words, this is case when $`0<1m1)`$. Then, we may set
$$A(x)A_{lim}a(x),$$
(20)
where $`a(x)`$ is positive but small and vanishes at $`x=+\mathrm{}`$. Using the smallness of $`a(x)`$ and $`B_1`$, one can simplify Eqs. (16) and (17):
$`(1/\sqrt{2})a^{\prime \prime }\left[\sqrt{2}(2\mu _2)aB_1^2\right]\mu _1B_1^2`$ $`=`$ $`0,`$ (21)
$`B_1^{\prime \prime }+\left[\sqrt{2}(2\mu _2)aB_1^2\right]B_1{\displaystyle \frac{1}{2}}\left(\mu _12\mu _2\right)B_1`$ $`=`$ $`0.`$ (22)
The b.c. for $`a(x)`$ at $`x=x_0`$ can be obtained from the expansion of the exact b.c. (19):
$$a(x_0)=\frac{B_1^2(x_0)}{\sqrt{2}\left(2\mu _2\right)}.$$
(23)
The lowest-order approximate analytical solution to Eqs. (21) and (22), satisfying the necessary boundary conditions, is very simple:
$$B_1(x)=B_1(x_0)\mathrm{exp}(\lambda x),a(x)=a(x_0)\mathrm{exp}(2\lambda x),$$
(24)
with $`\lambda =\sqrt{\left(\mu _22\mu _1\right)/2}`$. Note that the condition necessary for $`\lambda `$ to be real is again exactly tantamount to $`m<1`$.
## 4 Numerical Results
To check the qualitative predictions for the structure of the transient layer obtained in the previous section, we performed a two-stage numerical integration of Eqs. (14) and (15), and then of Eqs. (16) and (17). To this end, the first pair of the equations was solved with the b.c. (12), continuing the solution until it hits a point where it satisfies Eq. (19) (recall the latter condition is equivalent to the vanishing of $`B_2^2`$). Actually, the numerical integration was performed, instead of the original coordinate $`x`$, in terms of a variable $`\xi \mathrm{tanh}x`$. This transformation is convenient because it maps the semi-infinite intervals $`(\pm \mathrm{},0)`$ of the variable $`x`$ into finite ones $`(\pm 1,0)`$.
The values $`A(x_0)`$ and $`B_1(x_0)`$, obtained from the solution of Eqs. (14) and (15), where then used to find a solution to Eqs. (16) and (17) in the interval $`\mathrm{tanh}x_0<\xi <1`$, satisfying the b.c. (11) at $`\xi =1`$. In accord with what said above, a condition of the continuity of the first derivatives across the point $`\xi =\mathrm{tanh}x_0`$ was not imposed.
Three typical examples of the thus obtained numerical solutions are displayed vs. the coordinate $`x`$ in Fig. 2, for three characteristic values $`m=0.75`$, $`m=0.5`$, and $`m=0.25`$, the parameter $`\mu _2`$, that must be small enough, being fixed in all the three cases as $`\mu _2=0.1`$. A characteristic feature clearly seen in all the cases is that, in accord with the prediction of the above analysis, a width of the left sublayer is essentially larger than that of the right one. It is also noteworthy that the solutions changes very little with a large change in $`m`$, i.e., the transient layer between the two SC domains is expected to be quite robust.
## 5 Conclusion
In this work, we have proposed an approach that makes it possible to the generate a two-dimensional quasiperiodic pattern in nonlinear dissipative systems where a direct generation of stable uniform quasiperiodic planforms is not possible. An eightfold pattern can be created in the form of a broad transient stripe between two domains filled by square cells, which are oriented under the angle of $`45`$ degrees relative to each other. Using the symmetry of the configuration considered, the structure of the pattern was described in terms of a system of three coupled real stationary Ginzburg-Landau equations, which were analyzed by means of analytical and numerical methods. It was found that the transient quasiperiodic pattern exists exactly in a parametric region in which the uniform square-cell pattern is stable. Further, it was found that the transient layer consists of two different sublayers, with a derivative jump between them (that can be smoothed into an additional narrow boundary layer, if higher-order derivatives are added to the Ginzburg-Landau equations). The width of the sublayer that features the eightfold quasiperiodic pattern is found to be large, while the other sublayer (filled with a less interesting pattern, which is quasiperiodic only in one direction) has a width $`1`$. A broad stripe of a twelvefold QP pattern can be similarly generated as a transient layer between two domains of hexagonal cells oriented at the angle of $`30`$ degrees.
It still remains to perform simulations of the full time-dependent Ginzburg-Landau equations, in order to directly test the dynamical stability of the broad transient layers. However, the numerically found robustness of the layers against the variation of the crucial control parameter $`m`$ suggests that they have a good chance to be dynamically stable.
## Figure Captions
Fig. 1. The configuration giving rise to the transient layer filled with the eightfold quasi-periodic pattern between two domains of square cells oriented under the angle $`45`$ degrees relative to each other..
Fig. 2. Numerically found structure of the transient layer corresponding to the configuration shown in Fig. 1. The small parameter $`\mu _2`$ defined by Eq. (9) is fixed to be $`0.1`$, while the control parameter $`m`$, defined by Eq. (10), takes values $`0.75`$ (a), $`0.50`$ (b), and $`0.25`$ (c) (recall only the values $`0<m<1`$ make sense in the present context). |
warning/0003/astro-ph0003176.html | ar5iv | text | # Reionization Constraints on the Contribution of Primordial Compact Objects to Dark Matter
## 1 Introduction
Observations of the rotation curves of galaxies and clusters, in addition to joint fits of Type Ia supernova data and the power spectrum of the cosmic microwave background, suggest that the density of matter in the current universe is $``$30% of the closure density, i.e., $`\mathrm{\Omega }_m0.3`$. However, the success of big bang nucleosynthesis in explaining the primordial abundances of light elements, especially the primordial abundance ratio of D/H, requires that the contribution of baryons is only $`\mathrm{\Omega }_bh^2=0.019\pm 0.0024`$ (95% confidence; Tytler et al. 2000), where $`hH_0/100`$ km s<sup>-1</sup> Mpc<sup>-1</sup> and $`H_0`$ is the present day Hubble constant. The majority of the matter must be something else.
One class of possibilities involves hypothesized exotic particles, from light particles such as axions (Peccei & Quinn 1977) to heavier particles such as the neutralino (e.g., Jungman, Kamionkowski, & Greist 1996) or even ultramassive particles such as โWIMPZILLAsโ (Kolb, Chung, & Riotto 1998; Hui & Stewart 1999). Another class, which we focus on in this paper, involves dark matter that occurs primarily in $`0.11M_{}`$ clumps. This class, which has received recent attention because this is the mass scale of objects discovered by microlensing projects such as MACHO, EROS, and OGLE, has several specific candidates. For example, black holes may have formed during the QCD phase transition from quark matter to nucleonic matter (Jedamzik 1997, 1998; Niemeyer & Jedamzik 1999), during which the horizon mass was plausibly in the $`0.11M_{}`$ range. Other suggestions involve quark stars (Banerjee et al. 2000), boson stars (Colpi, Shapiro, & Wasserman 1986; Mielke & Schunck 2000), and stars formed of mirror matter (Mohapatra & Teplitz 1999). Here we consider those members of this class that involve primordial compact objects, specifically those objects which (1) existed before the $`z1100`$ epoch of decoupling, and (2) have a mass to radius ratio of $`GM/Rc^2\mathrm{ยฟ}\mathrm{}0.1`$. These include black holes, quark stars, and boson stars, but not mirror matter stars, as they are envisioned to form at comparatively late times and to be comparable to ordinary stars in their compactness (Mohapatra & Teplitz 1999).
Primordial compact objects will accrete from the ambient medium and will therefore generate substantial luminosity. This luminosity can ionize the surrounding medium. Unlike the energy spectra from ordinary stars, which drop off rapidly above the ground state ionization energy of hydrogen, the energy spectra from accreting compact objects are known observationally to be very hard, with substantial components above 1 keV and often extending above 100 keV. An important consequence of this is that whereas the Stromgren sphere of ionization around, say, an O or B star is extremely sharply defined, with an exponentially decreasing ionization fraction outside the critical radius, the ionization fraction produced by an accreting compact object dies off relatively slowly with radius, as $`r^{3/2}`$ (Silk 1971; Carr 1981). Therefore, accretion onto a primordial object can produce ionization over a large volume in the early universe. If the resulting optical depth to Thomson scattering is too large, it will conflict with the upper limit to this optical depth derived from the observed anisotropy of the microwave background (Griffiths, Barbosa, & Liddle 1999). Conversely, the upper limit on the optical depth can be used to constrain the properties of primordial compact objects, if these are proposed as the dominant component of dark matter.
Here we calculate the ionization produced by compact objects accreting in the early universe. We find that the ionization produced by secondary electrons, an effect not included in previous analyses of reionization by accretion, increases substantially the ionization fraction and hence the optical depth to Thomson scattering. In ยง 2 we show that the Thomson optical depth out to the $`z1100`$ redshift of decoupling is $`\tau 24[f_{\mathrm{CO}}ฯต_1(M/M_{})]^{1/2}(H_0/65)^1`$, where $`ฯต_1`$ is the accretion efficiency $`L/\dot{M}c^2`$ divided by 0.1 and $`f_{\mathrm{CO}}`$ is the fraction of matter in primordial compact objects. We compare this result to the current observational upper limit of $`\tau <0.4`$, and show that either low accretion efficiency or low mass is required if dark matter is mostly composed of primordial compact objects. In ยง 3 we consider low-efficiency accretion such as flows dominated by advection or wind outflow. We show that the constraints from ionization are especially tight on objects without horizons. In ยง 4 we place this result in the context of previous constraints on, for example, primordial black holes as the main component of dark matter. We also discuss future improvements to our result. In particular, we show that the expected accuracy of optical depth measurements with MAP and Planck could decrease the upper bound on $`f_{\mathrm{CO}}ฯต_1(M/M_{})`$ by a further factor of $``$100.
## 2 Calculation of Optical Depth
If the number density of baryons is $`n(z)z^3n_0`$ (where in this entire calculation we assume $`z1`$) and the ionization fraction is $`x(z)`$, then the optical depth to Thomson scattering between redshifts $`z`$ and $`z+dz`$ is
$$d\tau (z)=n(z)x(z)\sigma _Tds(z)$$
(1)
where $`\sigma _T=6.65\times 10^{25}`$ is the Thomson scattering cross section and
$$ds(z)=\frac{1}{H_0}\frac{cdz}{(1+z)E(z)}$$
(2)
is the distance traveled by a photon in this redshift interval. Here $`E(z)=\left[\mathrm{\Omega }_m(1+z)^3+\mathrm{\Omega }_R(1+z)^2+\mathrm{\Omega }_\mathrm{\Lambda }\right]^{1/2}`$ and $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_R`$, and $`\mathrm{\Omega }_\mathrm{\Lambda }`$ are the current contributions to the mass energy of the universe from, respectively, matter, curvature, and the cosmological constant. At $`z1`$ the first term dominates, so that $`E(z)\mathrm{\Omega }_m^{1/2}z^{3/2}`$ and $`ds(z)cH_0^1\mathrm{\Omega }_m^{1/2}z^{5/2}dz`$. The differential optical depth is then
$$d\tau (z)n_0z^3x(z)\sigma _Tds(z)=n_0x(z)\sigma _T\frac{c}{H_0\mathrm{\Omega }_m^{1/2}}z^{1/2}dz$$
(3)
(see also Haiman & Knox 1999). This needs to be integrated out to the $`z1100`$ redshift of decoupling to determine the optical depth to scattering in the early universe. The main unknown in this expression is the ionization fraction $`x(z)`$. In the remainder of this section, therefore, we compute the ionization produced by radiation from accreting compact objects. In ยง 2.1 we compute the luminosity and spectrum of this radiation. We assume Bondi-Hoyle accretion and a spectrum corresponding to that observed from many neutron stars and black hole candidates. In ยง 2.2 we use the ionization balance equation to calculate the ionization produced by a single source. We include the effects of ionization by secondary electrons, which is a significant effect not included in the analysis of pregalactic black hole accretion by Carr (1981). In ยง 2.3 we show that the ionizing flux from sources spread throughout the universe increases significantly the ionization fraction. Finally, in ยง 2.4 we calculate the optical depth to Thomson scattering out to the $`z1100`$ redshift of decoupling, including the effects of Compton cooling by the microwave background.
### 2.1 Luminosity and Spectrum of Radiation
Let us now consider accreting objects of mass $`M`$. Suppose that these masses are moving with the Hubble flow, so that the main parameter governing the accretion rate is the speed of sound in the gas at infinity, $`a_{\mathrm{}}=\sqrt{\mathrm{\Gamma }_1kT/\mu m_p}`$, where $`\mathrm{\Gamma }_1`$ is the polytropic index, $`m_p`$ is the mass of the proton, and $`\mu `$ is the mean molecular weight. For pure hydrogen ($`\mu =1/2`$), the mass accretion rate from a perfect gas with $`\mathrm{\Gamma }_1=5/3`$ is then
$$\dot{M}=1.2\times 10^{10}\left(\frac{M}{M_{}}\right)^2\left(\frac{\rho _{\mathrm{}}}{10^{24}\mathrm{g}\mathrm{cm}^3}\right)T_4^{3/2}\mathrm{g}\mathrm{s}^1,$$
(4)
where $`T_4T_{\mathrm{}}/10^4`$ K and $`T_{\mathrm{}}`$ is the temperature of the gas at infinity. For a primordial composition of 75% hydrogen and 25% helium by mass, this accretion rate is more than doubled because helium has half the velocity for a given temperature that hydrogen does, and hence accretes at eight times the rate for a given mass density. We therefore take the coefficient to be $`3\times 10^{10}`$. If accretion produces a luminosity with an efficiency $`0.1ฯต_1`$, so that $`L=10^{20}ฯต_1\dot{M}`$ erg s<sup>-1</sup>, then
$$L=3\times 10^{30}\left(\frac{M}{M_{}}\right)^2\left(\frac{\rho _{\mathrm{}}}{10^{24}\mathrm{g}\mathrm{cm}^3}\right)T_4^{3/2}\mathrm{erg}\mathrm{s}^1.$$
(5)
The best estimate of the baryon density in the current universe from big bang nucleosynthesis constraints (Tytler et al. 2000) is
$$\rho _{B0}=3.6\pm 0.4\times 10^{31}\mathrm{g}\mathrm{cm}^3.$$
(6)
At a redshift $`z`$ this density is therefore $`(1+z)^3\rho _{B0}z^3\rho _{B0}`$. Hence, if the compact object accretes matter with the average baryonic density in the universe, the luminosity at redshift $`z`$ is
$$L10^{24}z^3\left(\frac{M}{M_{}}\right)^2T_4^{3/2}\mathrm{erg}\mathrm{s}^1.$$
(7)
Pressure balance of a hot HII region with the cooler exterior universe may decrease the density of accreting matter and therefore decrease this luminosity (see below). In accreting black hole sources from stellar mass to AGN, and also in some accreting neutron stars, the spectrum often has a power-law tail with equal power in equal logarithmic intervals of the photon energy, up to some $`E_{\mathrm{max}}`$: $`dL(E)/dEE^1\mathrm{exp}(E/E_{\mathrm{max}})`$. The results of our calculation are fairly insensitive to the assumed spectrum. Normalizing this spectrum so that the total luminosity above $`E_0=13.6`$ eV is $`L`$, the differential photon flux at energy $`E`$ a distance $`R`$ from the compact object is
$$F(E)=e^{\tau (E)}\frac{dL/dE}{4\pi R^2E}=e^{\tau (E)}\frac{L}{4\pi \mathrm{ln}(E_{\mathrm{max}}/E_0)E^2R^2}e^{E/E_{\mathrm{max}}}.$$
(8)
Here $`\tau (E)`$ is the optical depth at a distance $`R`$ from the source to photons of energy $`E`$. Note that Carr (1981) chose a spectrum of a bremsstrahlung form ($`dL/dEe^{E/E_{\mathrm{max}}}`$), and hence had a different energy dependence and normalization for the photon number flux.
### 2.2 Secondary Ionization and Ionization Balance
A given photon can effectively produce many ionizations, because the ionized electrons can collisionally ionize other atoms (see, e.g., Silk & Werner 1969; Silk 1971). The collisional cross section exceeds $`10^{17}`$ cm<sup>2</sup> for electron energies between $``$15 eV and 1 keV (see Dalgarno, Yan, & Liu 1999 for a recent discussion of electron energy deposition). Dalgarno et al. (1999) calculate that the mean energy per ion pair decreases with increasing initial electron energy, reaching a limit of 36.1 eV per pair at energies $`>`$200 eV. Using their Figure 6, we adopt an approximate value of $`E/3E_0`$ hydrogen atoms ionized by a photon of initial energy $`E`$; this is a rough average over the energy range of interest, and we assume for simplicity that it is constant over that range.
The effective ionization rate produced by the photons generated by accretion is therefore (adapting the formula of Carr 1981)
$$\zeta _H_{E_0}^{\mathrm{}}\sigma _1\left(\frac{E}{E_0}\right)^3\left(\frac{E}{3E_0}\right)F(E)๐E,$$
(9)
where $`\sigma _12\times 10^{17}`$ cm<sup>2</sup> is the abundance-weighted ionization cross section at $`E_0`$. The integrand in this formula is a constant factor
$$\frac{E_{\mathrm{max}}}{3E_0\mathrm{ln}(E_{\mathrm{max}}/E_0)}$$
(10)
times the integrand in the corresponding formula in Carr (1981). The difference arises because we assume a different form for the spectrum and account for the ionization produced by secondary electrons. The remainder of the analysis of the ionization region created by a single source follows the treatment of Carr (1981), with this factor included. This is a large factor, of order 25 for $`E_{\mathrm{max}}=10^8`$ erg, and it therefore makes a crucial difference to the overall ionization. If $`E_{\mathrm{max}}E_0`$, the ionization rate is approximately (Silk et al. 1972)
$$\zeta _H\left(\frac{E_{\mathrm{max}}}{3E_0\mathrm{ln}(E_{\mathrm{max}}/E_0)}\right)\frac{\sigma _1L}{12\pi E_{\mathrm{max}}R^2}\left[\frac{1\mathrm{exp}(\tau _0)}{\tau _0}\right],$$
(11)
where
$$\tau _0=_0^Rn_H(1x)\sigma _1๐R$$
(12)
and the first factor in parentheses indicates the correction factor to the expression of Carr (1981). Here $`x`$ is the ionized fraction at radius $`R`$.
The ionization balance equation is
$$\alpha n_H^2x^2=\zeta _Hn_H(1x),$$
(13)
where around $`T10^4`$ K, the recombination coefficient not including single-photon transitions to the ground state (which would release ionizing photons) is $`\alpha 2.6\times 10^{13}T_4^{0.75}`$ cm<sup>3</sup> s<sup>-1</sup> (Hummer 1994). Far from the accreting compact object, where $`x1`$ and $`\tau _01`$, the ionization fraction is
$$x=\left(\frac{E_{\mathrm{max}}}{3E_0\mathrm{ln}(E_{\mathrm{max}}/E_0)}\right)^{1/2}\frac{1}{\sqrt{8}}\left(\frac{R}{R_s}\right)^{3/2},$$
(14)
where
$$R_s=\left[\frac{2L}{3\pi \alpha n_H^2E_{\mathrm{max}}}\right]^{1/3}.$$
(15)
Again, the initial factor in the equation for $`x`$ is the correction factor, which therefore increases the ionization fraction far from the compact object by a factor of $``$5; note that the only remaining dependence on $`E_{\mathrm{max}}`$ is $`\left[\mathrm{ln}(E_{\mathrm{max}}/E_0)\right]^{1/2}`$, so this result is very insensitive to the high-energy cutoff of the spectrum.
### 2.3 Contribution of Multiple Sources
The total ionization rate $`\zeta _H`$ must be summed over the contributions of all sources. At large distances from a source, $`\tau _0R1`$, so that $`\zeta _HR^3`$. For multiple sources separated by an average distance $`R_{\mathrm{sep}}`$, the ionizing rate is larger than the single-source ionizing rate at a distance $`R_{\mathrm{sep}}`$ by a factor
$$\frac{\zeta _H}{\zeta _H(r=R_{\mathrm{sep}})}=_{R_{\mathrm{sep}}}^{R_{\mathrm{max}}}\left(\frac{r}{R_{\mathrm{sep}}}\right)^34\pi r^2n_{\mathrm{CO}}๐r.$$
(16)
Here $`n_{\mathrm{CO}}=10^{62}z^3(M/M_{})^1\mathrm{\Omega }_{\mathrm{CO}}`$ cm<sup>-3</sup> is the number density of compact objects at redshift $`z`$, where $`\mathrm{\Omega }_{\mathrm{CO}}`$ is the fraction of the closure density in compact objects. Also, $`R_{\mathrm{max}}\mathrm{min}[10^{31}z^3x^1,c/H(z)]`$ cm is the mean free path to Thomson scattering. The separation distance is approximately given by $`\left(\frac{4}{3}\pi R_{\mathrm{sep}}^3\right)^1=n_{\mathrm{CO}}`$, so $`R_{\mathrm{sep}}^3\frac{3}{4\pi }n_{\mathrm{CO}}^1`$. Therefore,
$$\frac{\zeta _H}{\zeta _H(r=R_{\mathrm{sep}})}3\mathrm{ln}(R_{\mathrm{max}}/R_{\mathrm{sep}}).$$
(17)
The ratio of radii is typically $`10^610^8`$, so the enhancement due to the contributions of multiple sources is approximately a factor of 50.
When multiple sources are included, the ionization fraction (for $`x1`$) is increased by a factor that is approximately the square root of the factor by which the ionization rate is enhanced. At a distance $`R`$ the rate is enhanced by a factor
$$\zeta _H\zeta _H\left(1+50(R/R_{\mathrm{sep}})^3\right),$$
(18)
and hence the ionization fraction including multiple sources is
$$x\left(\frac{E_{\mathrm{max}}}{3E_0\mathrm{ln}(E_{\mathrm{max}}/E_0)}\right)^{1/2}\frac{1}{\sqrt{8}}\left(\frac{R}{R_s}\right)^{3/2}\left(1+50(R/R_{\mathrm{sep}})^3\right)^{1/2}.$$
(19)
Integrating this from $`R_s`$ to $`R_{\mathrm{sep}}`$, the volume-averaged ionization is
$$\overline{x}3\left(\frac{E_{\mathrm{max}}}{3E_0\mathrm{ln}(E_{\mathrm{max}}/E_0)}\right)^{1/2}\left(\frac{R_s}{R_{\mathrm{sep}}}\right)^{3/2}.$$
(20)
Here $`R_{\mathrm{sep}}=3\times 10^{20}z^1(M/M_{})^{1/3}\mathrm{\Omega }_{\mathrm{CO}}^{1/3}`$ cm.
### 2.4 Optical Depth Including Compton Cooling and Pressure Balance
An evaluation of this expression for the ionization fraction requires knowledge of the luminosity of individual sources and the average temperature of the matter in the universe. As pointed out by, e.g., Carr (1981), the dominant cooling process at high redshifts is inverse Compton cooling off of the microwave background. If the temperature $`T`$ of the matter is much larger than the temperature $`T_r`$ of the radiation background, then the cooling rate per volume at redshift $`z`$ is
$$\mathrm{\Gamma }_r2\times 10^{38}x(z)z^7T_4\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1.$$
(21)
The average heating rate is just the luminosity per source times the number density of sources:
$$\mathrm{\Gamma }_h=Ln_{\mathrm{CO}}=10^{38}\mathrm{\Omega }_{\mathrm{CO}}z^6\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1.$$
(22)
At $`z1000`$, where the optical depth to Thomson scattering exceeds unity and as we will see $`x0.1`$, the cooling rate dominates the heating rate and hence the matter temperature is locked to the radiation temperature during this epoch (see also Carr 1981). This increases the recombination rate over most of the volume of interest, and therefore decreases the optical depth to scattering. Inside the HII region, by contrast, heating dominates cooling and the temperature remains close to $`10^4`$ K for $`z>10`$; in fact, Carr (1981) finds that the temperature is $`T_4=(z/10^3)^{0.3}`$. The temperature difference means that pressure balance requires that the density inside the HII region be less than the average density by a factor $`\overline{T}/T`$; note, however, that this configuration requires the support against gravity of a denser by a less dense medium, which therefore is in principle Rayleigh-Taylor unstable. Hence, mixing could occur which would decrease the temperature and increase the density of the HII region. If mixing does not occur, the density of the matter accreting onto the compact object is decreased by a factor $`0.27(z/10^3)^{0.7}`$. The ionization fraction and hence the optical depth would therefore be reduced by the square root of this factor, or about $`0.5(z/10^3)^{0.35}`$. The uncertainty of whether there is an interchange instability and mixing thus produces an uncertainty of a factor $``$2 in the optical depth to scattering.
With these contributions, the average ionization is
$$\overline{x}(z)510\times 10^4ฯต_1^{1/2}\left(\frac{M}{M_{}}\right)^{1/2}\mathrm{\Omega }_{\mathrm{CO}}^{1/2}\left[\frac{6}{\mathrm{ln}(E_{\mathrm{max}}/E_0}\right]^{1/2}z^{0.7}.$$
(23)
The differential optical depth to scattering is $`d\tau (z)=n_0x(z)\sigma _T\frac{c}{H_0\mathrm{\Omega }_m^{1/2}}z^{1/2}dz`$, so using $`n_0=2.2\pm 0.3\times 10^7`$ cm<sup>-3</sup> and evaluating the constants, this is
$$d\tau (z)12\times 10^6\left[\frac{6}{\mathrm{ln}(E_{\mathrm{max}}/E_0)}\right]^{1/2}ฯต_1^{1/2}(M/M_{})^{1/2}(H_0/65\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1)^1f_{\mathrm{CO}}^{1/2}z^{1.2}dz.$$
(24)
Here $`f_{\mathrm{CO}}\mathrm{\Omega }_{\mathrm{CO}}/\mathrm{\Omega }_m`$ is the fraction of matter in compact objects. Integrating from a small redshift to the redshift $`z1100`$ at decoupling gives finally
$$\tau 24\left[\frac{6ฯต_1}{\mathrm{ln}(E_{\mathrm{max}}/E_0)}\left(\frac{M}{M_{}}\right)f_{\mathrm{CO}}\right]^{1/2}\left(\frac{H_0}{65}\right)^1.$$
(25)
The effect of this optical depth on the observed CMB power spectrum is not identical to the effect of the same optical depth if it came from sudden and complete reionization at some lower redshift $`z1040`$. The reason is that the mechanism described here produces most of the optical depth at comparatively high redshifts, $`z\mathrm{ยฟ}\mathrm{}800`$, and hence for optical depths in excess of unity the scatterings occur close to recombination where some of the primordial anisotropy is maintained. In contrast, scattering at low redshift exponentially suppresses the primordial anisotropy. However, if the optical depth is less than unity this effect is less pronounced in the reionization mechanism discussed in this paper, because scatterings occur over a wide range of redshift and hence tend to smooth out small-scale anisotropies in the same way as would happen due to scattering at much lower redshifts. These qualitative effects are confirmed by simulations with CMBFAST (Seljak & Zaldarriaga 1996 and subsequent papers), which show that for $`\tau \mathrm{ยก}\mathrm{}1`$ the constraints on the optical depth from the observed CMB power spectrum are roughly the same for this mechanism as for late reionization.
The observational upper limit to $`\tau `$ from small-scale CMB anisotropy is $`\tau \mathrm{ยก}\mathrm{}0.4`$ if $`\mathrm{\Omega }_m=0.3`$ and the primordial power spectrum has an index $`n=1`$ (Griffiths et al. 1999). To be consistent with this limit, primordial compact objects must therefore be either low-efficiency accretors, low-mass objects, or a minor component of dark matter. Given that the measured mass spectrum of MACHOs in our galaxy has a peak in the $`0.5M_{}1M_{}`$ range (Alcock et al. 2000; note, however, that most of the mass in the halo need not be in MACHOs \[Gates et al. 1998, Alcock et al. 2000\] and a higher-mass component is not ruled out \[Lasserre et al. 2000\]), explanation of these objects with a population that composes most of the dark matter in the universe requires low-efficiency accretion, which we consider in the next section. These constraints are particularly strict for higher-mass black holes. The joint limits on $`M`$ and $`ฯต`$ are shown in Figure 1, for three different upper limits to the Thomson optical depth: $`\tau `$=0.4, 0.14, or 0.05, which are the optical depths obtained if the the universe was fully and suddenly reionized at a redshift of $`z_{\mathrm{reion}}`$=40, 20, or 10, respectively.
## 3 Efficiency of Accretion
In the last few years there has been much discussion of the possibility that, for low accretion rates, an advection-dominated accretion flow (ADAF) is set up in which the radiative efficiency onto black holes is low because the matter flows almost radially into the hole, taking almost all of its energy with it (e.g., Narayan & Yi 1994; Abramowicz et al. 1995). The radiative efficiency according to these solutions could be extremely low, and hence might allow a large matter density in black holes. If the compact object does not have a horizon, then an ADAF will not reduce the radiative efficiency, so this is not a way out.
Another possibility is that most of the accreting matter does not reach the surface at all, perhaps because it is driven out in a wind. This is the basis of the advection-dominated inflow-outflow solution (ADIOS) proposed by Blandford & Begelman (1999). Convective flow, in which the net inward flow rate at small radii is small, has also been found analytically by Quataert & Narayan (1999) and in numerical simulations by Stone, Pringle, & Begelman (1999). If in Bondi-Hoyle accretion the result of the accretion is inflow of a small fraction of matter combined with outflow of most of the matter, then the accretion efficiency onto even objects without horizons could be small. However, there is a crucial unsolved problem with these flows, which is whether they will remain at low efficiency indefinitely if there is a steady inflow of matter from infinity, as in Bondi-Hoyle accretion. If instead the accretion proceeds as in a dwarf nova, in which there is a long-term buildup of matter followed by a short-term, high-luminosity episode during which the accumulated matter is dumped onto the central object, then current accretion theory suggests the accretion will generate radiation efficiently regardless of the nature of the compact object. In such a case, neither black holes nor any other type of primordial compact object are viable candidates for most of the dark matter in the universe.
## 4 Discussion and Conclusions
Consideration of compact objects as components of dark matter has often been restricted to black holes, but many of the arguments apply more generally. Black holes with masses in excess of $`10^3M_{}`$ are ruled out as a significant component of galactic halos because their dynamical interactions with globular clusters would destroy the clusters (for a recent calculation see Arras & Wasserman 1999). The lack of an increase in the number of low equivalent width quasars with increasing redshift (expected to be caused by gravitational lensing) rules out a contribution $`\mathrm{\Omega }\mathrm{ยฟ}\mathrm{}0.1`$ from any objects with masses between $`10^2M_{}`$ and $`20M_{}`$ that are more compact than their Einstein radii (Dalcanton et al. 1994). The lack of observed lensing of cosmological gamma-ray bursts also allows weak limits to be placed on the contribution of black holes of various sizes: $`\mathrm{\Omega }<0.15`$ at the 90% level for $`M=10^{6.5}M_{}`$, $`\mathrm{\Omega }<0.9`$ at the $`1\sigma `$ level for $`M=10^{12.5}10^9M_{}`$, and $`\mathrm{\Omega }<0.1`$ ($`z_{\mathrm{GRB}}1`$) or $`\mathrm{\Omega }<0.2`$ ($`z_{\mathrm{GRB}}2`$) at the 95% level for $`M=10^{16}10^{13}M_{}`$ (Marani et al. 1999).
Here we show that ionization from compact object accretion in the early post-decoupling universe is more significant than had been thought previously, because of the effects of secondary ionization by electrons. The result is that, barring inefficient accretion ($`ฯต<0.05`$ for $`M=0.1M_{}`$, $`ฯต<0.005`$ for $`M=1M_{}`$), primordial compact objects in this mass range cannot compose a significant fraction of the mass of the universe, because they would ionize the universe enough to conflict with the measured small-scale anisotropies of the cosmic microwave background. If further analysis and numerical simulation of flows onto black holes demonstrates that the long-term time averaged accretion efficiency is $`\mathrm{ยฟ}\mathrm{}0.1`$, as might happen if matter tends to pile up as in a dwarf nova and then accrete quickly with efficient radiation, then all masses greater than $`0.1M_{}`$ are excluded from making a significant contribution.
Future CMB missions such as MAP and Planck could strengthen these constraints considerably. The optical depth resolution of MAP is expected to be 0.022, and of Planck is expected to be 0.004 (Zaldarriaga, Spergel, & Seljak 1997; Bouchet, Prunet, & Sethi 1999; Eisenstein, Hu, & Tegmark 1999). Since the existence of a Ly$`\alpha `$ emitter at $`z=5.64`$ (see Haiman & Spaans 1999) shows that reionization must have occurred before then, this means that both satellites, and especially Planck, will be able to detect the effects of ionization regardless of the actual redshift of reionization. If $`z_{\mathrm{reion}}10`$ then the redshift of reionization could even be determined directly with SIRTF or NGST via, e.g., analysis of the damping wing of the Gunn-Peterson trough (Miralda-Escudรฉ 1998) or detection of transmitted flux between Lyman resonances (Haiman & Loeb 1999). The upper limit on the product $`ฯต_1(M/M_{})f_{\mathrm{CO}}`$ scales like $`\tau _{\mathrm{scatt}}^2`$ (or, for $`z1`$, like $`z_{\mathrm{reion}}^3`$), so if $`z_{\mathrm{reion}}10`$ this upper limit is decreased by almost a factor of 100. In this case, barring extremely inefficient accretion, dark matter must be composed of less compact objects or of WIMPs.
We thank Sylvain Veilleux, Andy Young, Jim Stone, Eve Ostriker, and Scott Dodelson for comments. This work was supported in part by NASA ATP grant number NRA-98-03-ATP-028. |
warning/0003/gr-qc0003069.html | ar5iv | text | # Gravitational RadiationTo be published in the Encyclopedia of Astronomy and Astrophysics (Institute of Physics Publishing, Bristol, and Macmillan Publishers Ltd, London, 2000).
## Gravitational wave astronomy
Gravity is one of the fundamental forces of Nature, and it is the dominant force in most astronomical systems. In common with all other phenomena, gravity must obey the principles of Special Relativity. In particular, gravitational forces must not be transmitted or communicated faster than light. This means that when the gravitational field of an object changes, the changes ripple outwards through space and take a finite time to reach other objects. These ripples are called gravitational radiation or gravitational waves.<sup>1</sup><sup>1</sup>1They are also sometimes referred to as gravity waves, but since this term has a different meaning in meteorology and stellar hydrodynamics, we will avoid it here. See Solar Interior: Helioseismology.
In Einsteinโs theory of gravitation (see General Relativity and Gravitation), as in many other modern theories of gravity (see Non-general Relativity Theories of Gravity), gravitational waves travel at exactly the speed of light. Different theories make different predictions, however, about details, such as their strength and polarization. There is strong indirect observational evidence (see Binary Stars as a Probe of General Relativity, Hulse-Taylor Pulsar) that gravitational waves follow the predictions of general relativity, and instruments now under construction are expected to make the first direct detections of them in the first years of the 21st century.
These instruments and plans for future instruments in space are described in the article Gravitational Radiation Detection on Earth and in Space. Detectors must look for gravitational radiation from astronomical systems, because it is not possible to generate detectable levels of radiation in the laboratory. It follows that gravitational wave detection is also a branch of observational astronomy.
The most striking aspect of gravitational waves is their weakness. A comparison with the energy in light will illustrate this. The human eye has no trouble sensing the light from the planet Jupiter: the amount of energy that passes through the iris of the eye is far more than the minimum the eye can detect. Yet several times a week a gravitational wave, generated in a far distant galaxy, carries a similar amount of energy into the eye, and we donโt notice it.
The reason is that gravity is the weakest of the fundamental forces, and the disturbance created by even such an energetic wave is so tiny that no man-made instrument has so far registered it. While all the energy in the light from Jupiter that enters the eye is absorbed in the eye, the gravitational wave passes right through, leaving behind almost none of its energy. All the matter in the present Universe is similarly transparent to gravitational waves.
Gravitational radiation is today one of the last unopened windows into the Universe. There are at least five reasons motivating scientists to develop gravitational wave astronomy:
* The weakness with which gravitational waves interact with matter is a great advantage for astronomy. It means that gravitational waves arrive unaffected by any intervening matter they may have encountered since being generated. There is no significant scattering or absorption, although they will be deflected by a Gravitational Lens in the same way as light. Gravitational waves carry uncorrupted information even if they come from the most distant parts of the Universe or from its most hidden regions, like the interiors of Supernovae.
* Gravitational waves are emitted by the bulk motions of their sources, not by individual atoms or electrons, as is normally the case for electromagnetic waves. They therefore carry a completely different kind of information about their sources from that which is normally available in observations of Binary Stars, supernovae, and Neutron Stars.
* Gravitational waves can be emitted by black holes, which are described in the article General Relativity and Gravitation. Indeed, gravitational waves provide only way to make direct observations of these objects. Since there is now strong indirect evidence that giant black holes inhabit the centers of many (or even most) galaxies (see Supermassive black holes in AGN), and since smaller ones are common in the Galaxy (see Black Hole Candidates in X-Ray Binaries), there is great interest in making direct observations of them.
* Gravitational waves can come from extraordinarily early in the history of the Universe. The electromagnetic radiation from the Big Bang is called the Cosmic Microwave Background. Observations of it describe the Universe at it was about $`10^5`$ years after the Big Bang. Studies of cosmological Nucleosynthesis give information about what the Universe was like as little as 3 minutes after the Big Bang. Gravitational waves, if they can be detected, would picture the Universe when it was only perhaps $`10^{24}`$ seconds old, just at the end of Inflation.
* Gravitational radiation is the last fundamental prediction of Einsteinโs general relativity that has not yet been directly verified. If another theory of gravity is correct, then differences could in principle show up in the properties of gravitational waves, such as their polarization. In principle, there must be a better theory of gravity, since general relativity is not a quantum theory, a deficiency that theoretical physicists today are working hard to remedy. The majority belief today is that there should be a unified theory of the fundamental forces, in which gravitation is related to the other forces. Evidence for the nature of this relation could show up in observations of gravitational waves, particularly those from the Big Bang.
These motivations and their implications are developed in the following sections. Each section begins with an introduction to the physical ideas and then develops some of the mathematical details.
## The physics of gravitational radiation: weakness and strength
The starting point for understanding gravitational radiation is Newtonian gravity. The weakness of gravity is evident. If a child picks up a book, she defeats the cumulative gravitational pull of the entire planet Earth on the book. The strength to do this comes from the chemical forces in her muscles, which come from electromagnetic interactions.
In fact, the electromagnetic force between the electron and the proton in a hydrogen atom is $`2\times 10^{39}`$ times bigger than the gravitational force between them. The reason that gravity can nevertheless dominate on the cosmic scale is that opposite electrical charges cancel each other, while the gravitational forces of all the particles add.
Another fact about gravity that was known to Newton is what is now called the equivalence principle (see General Relativity and Gravitation). This is the principle that all bodies accelerate in the same way in a gravitational field, so that the trajectory that a freely falling body (a body influenced only by gravity) follows in a given gravitational field depends only on its starting position and velocity, not on what it is made of.
Imagine now a machine made in some way to detect gravitational waves. Whatever the method of detection, a wave needs somehow to alter the internal state of the detector. If the wave carries a gravitational field that is completely uniform across the detector, then by the equivalence principle all of the parts of the detector will accelerate together, and its state will not change at all. To detect a gravitational wave, the machine must measure the non-uniformities of the gravitational field across a detector.
These non-uniformities are called tidal forces, because they produce the stretching effects that raise tides on the Earth. Gravitational waves are traveling tidal forces.
Newtonโs theory of gravity had no gravitational waves. For Newton, if a gravitational field changed in some way, that change took place instantaneously everywhere in space. This is not a wave. Let us consider what we mean by the term โwaveโ in ordinary language.
Imagine a childโs rubber duck floating in a bath tub half full of water. If a child presses down on the duck very gently, until is is nearly submerged, then the level of the water will rise everywhere in a nearly uniform way, and this is not called a wave. If instead he drops the duck, then the disturbance rises around the base of the duck rapidly, moves away from it, and eventually reaches the walls. This is a wave. Wave motion requires a finite speed for the propagation of disturbances. If the disturbance is very slow, as for the floating duck, then the wavelength is very long, and near the site of the disturbance the wave motion is not noticeable. We say we are in the โnear zoneโ. But when we are more than a wavelength away, then we see waves, and this is the โwave zoneโ or โfar zoneโ. For the dropped duck, we see that waves because their wavelengths are shorter than the size of the bath. In general relativity, the speed of gravity is the speed of light. Because of this finite speed, gravity must exhibit wave effects.
Many of Newtonโs contemporaries were unhappy that his theory of gravity was based on instantaneous โaction at a distanceโ, but Newtonโs theory fit the observational facts. If gravity had a finite speed of propagation, there was no evidence for it in the solar system. Interestingly, the brilliant 18th century French mathematician and physicist Laplace tried out a variation on Newtonโs theory in which gravity was represented by something โflowingโ out of its source with a finite speed. He reasoned that a planet like the Earth, moving through this fluid of gravity, would experience friction and gradually spiral in towards the Sun.
Laplace could show that the observational limits on this inspiral even in his day were so stringent that the speed of gravity in his model needed to be huge compared to the speed of light. He did not find this result attractive and took the theory no further. (Laplace also explored the notion of what we now call a black hole, which for him was a region where gravity was strong enough to trap light.)
It is interesting that today, observations of the two neutron stars in the binary system PSR1913+16 spiraling together as they orbit one another provides the most convincing evidence that gravitational waves exist and are as described in general relativity. (See below and the article Binary Stars as a Probe of General Relativity.) Laplace had the right effect, but the wrong theory. This evidence is described in the next section.
In general relativity, Einstein used the principle of equivalence as the basis for a geometrical description of gravity. In the four-dimensional world of space-time, the trajectory of a particle falling freely in a gravitational field is a certain fixed curve. Its direction at any point depends on the velocity of the particle. The equivalence principle implies that there is a preferred set of curves in space-time: at any point, pick any direction, and there is a unique curve in that direction that will be the trajectory of any particle starting with that velocity. These trajectories are thus properties of space-time itself.
Moreover, if there were no gravitational field, the trajectories would be simple straight lines. Even in a gravitational field, a small freely falling particle does not โfeelโ any acceleration: its internal state is the same as if there were no gravity. Therefore Einstein postulated that a gravitational field made space-time curved, and that the preferred trajectories were locally straight lines that simply changed direction as they moved through the curved space-time, in much the same way as a great circle on a sphere changes direction relative to other great circles as one goes along it. For weak gravitational fields of slowly moving bodies, Einsteinโs theory reduces to Newtonโs in the first approximation.
For gravitational waves, one could make a very simple detector just by monitoring the distance between two nearby freely falling particles. If they are genuinely free, then any changes in their separations would indicate the passage of a gravitational wave. Because this measures a tidal effect, the bigger the separation of the particles, the bigger will be the change in their separation, at least for particles that are separated by less than a gravitational wavelength. Most modern gravitational wave detectors are designed to be as big as cost and practicality allows. These are described in the article Gravitational Radiation Detection on Earth and in Space.
Although gravitational radiation is well understood in theoretical terms in general relativity, the complexity and non-linearity of Einsteinโs equations means that calculations are often difficult. In the historical development of general relativity, between 1915 and the 1950โs and 1960โs, these mathematical difficulties created confusion over the physical nature of gravitational radiation, and in particular over whether they carried energy away from the source. Improved mathematical techniques finally resolved the matter in favor of the simple physical picture presented here, but this picture would not be complete without the strong mathematical underpinning that now exists.
The question of energy in gravitational waves is still a delicate one. There is no question that waves carry energy (and momentum) away from their sources. Nevertheless, it is not possible in general relativity to localize the energy in the radiation to regions smaller than about a wavelength. Indeed the equivalence principle shows that โpointโ particles feel nothing, no matter how strong the wave. The wave only acts by stretching space-time, producing a tidal distortion in the separations between particles (see the discussion of polarization below).
For this reason, energy is localized only in regions, not at points. It is nevertheless real energy: the nonlinearity of general relativity allows waves to create gravitation themselves. Recent numerical simulations have shown that focussed gravitational waves can actually form black holes, trapping themselves. If the waves are weak, they enter the focussing region and re-emerge. If they are strong enough, they enter and never leave.
### Gravitational waves in a quasi-Newtonian model
It is possible to calculate the approximate size of the effect of a given gravitational wave by beginning with Newtonian gravity and adding waves to it. In Newtonian gravity the gravitational field produced by a mass $`M`$ at a distance $`r`$ is given by
$$\varphi =GM/r,$$
(1)
where $`G`$ is Newtonโs gravitational constant. The field of a gravitational wave must be a ripple on this, which means a small change that oscillates in space and time. A suitable form for a change that propagates at the speed of light in the $`z`$-direction with an angular frequency $`\omega `$ is:
$$\delta \varphi =ฯต\frac{GM}{r}\mathrm{sin}[\omega (z/ct)],$$
(2)
where $`ฯต`$ is a dimensionless number that would be expected to be small compared to 1. Its size is the subject of the next main section.
The field $`\delta \varphi `$ produces an acceleration in the $`z`$-direction that depends on its $`z`$-derivative. Both $`1/r`$ and the $`\mathrm{sin}()`$ term depend on z. The derivative of $`1/r`$ will be proportional to $`1/r^2`$, which is how the acceleration falls off in Newtonโs theory (where $`\varphi `$ is the only field). But the derivative of the $`\mathrm{sin}()`$ term does not change the $`1/r`$; rather, it essentially just multiplies $`\delta \varphi `$ by $`\omega /c`$. At sufficiently large distances from the source, this term will dominate the $`1/r^2`$ term and the acceleration produced by the wave will be:
$$\delta a_z=ฯต\omega \frac{GM}{rc}\mathrm{cos}[\omega (z/ct)].$$
(3)
Note that this term would not be present in the $`x`$\- and $`y`$-derivatives, so these components of the acceleration are much smaller in this quasi-Newtonian model of a gravitational wave.
### Effect on a simple detector
The tidal part of this acceleration, for a detector that has size $`\mathrm{}`$ in the $`z`$-direction, is to a first approximation
$$\mathrm{}\frac{d}{dz}\delta a_z=ฯต\omega ^2\mathrm{}\frac{GM}{rc^2}\mathrm{sin}[\omega (z/ct)].$$
(4)
If a detector consists of two freely falling particles with this relative acceleration, the equation of motion for their separation $`\mathrm{}`$ will be
$$\frac{d^2\mathrm{}}{dt^2}=ฯต\omega ^2\mathrm{}\frac{GM}{rc^2}\mathrm{sin}[\omega (z/ct)],$$
(5)
The dimensionless coefficient $`ฯต(GM/rc^2)`$ is typically very small. Even if $`ฯต`$ is of order 1, the other number is, with reasonable values for $`M`$ and $`R`$,
$$\frac{GM}{rc^2}=2.4\times 10^{21}\left(\frac{M}{M_{}}\right)\left(\frac{r}{20\mathrm{Mpc}}\right)^1.$$
The mass and distance scales here are those appropriate to a neutron star in the nearest large cluster of galaxies, the Virgo Cluster. (The distance unit is based on the astronomersโ parsec, denoted pc, which is about $`3\times 10^{16}`$ m. The unit Mpc is a megaparsec.) It is believed that several neutron stars are formed in the Virgo Cluster each year in supernova explosions. Ever since the beginning of the development of gravitational wave detectors such events have been high on the list of possible sources of gravitational waves.
To solve Equation (5), one takes $`r`$ and $`z`$ as constants, and uses the smallness of the right-hand-side, which implies that the changes in $`\mathrm{}`$ are tiny compared to $`\mathrm{}`$ itself. On the right-hand side one can therefore replace $`\mathrm{}`$ by $`\mathrm{}_0`$, the initial value of $`\mathrm{}`$, and then simply integrate twice in time to get (for an initial value $`d\mathrm{}/dt=0`$)
$$\frac{\mathrm{}(t)\mathrm{}_0}{\mathrm{}_0}=ฯต\frac{GM}{rc^2}\mathrm{sin}[\omega (z/ct)].$$
(6)
The right-hand-side of Equation (6) is identical to that of Equation (2). This is an important conclusion which fits neatly with Einsteinโs geometrical conception of gravity: the size of a gravitational wave gives directly the stretching of the distance between nearby free particles. It is conventional to call this $`h/2`$ and refer to $`h`$ as the gravitational wave potential
$$h:=2\frac{\mathrm{}(t)\mathrm{}_0}{\mathrm{}_0}=2\delta \varphi .$$
(7)
The amplitude of the oscillations of $`h`$ is
$$h2ฯต\frac{GM}{rc^2}.$$
(8)
It is evident from this that a detector must be able to measure changes in its own size that are smaller than one part in $`10^{21}`$ to have a reasonable chance of making astronomical observations. The extraordinary smallness of this effect also explains why ordinary objects in the Universe are transparent to gravitational waves. As the waves pass through them, they disturb them so little (parts per $`10^{21}`$ typically) that the transfer of energy to the object and any back-reaction effects of this on the wave are negligibly small.
### Energy flux carried by waves
The energy in the waves can also be estimated from these equations and general physical principles. Quite generally, in classical field theories, the energy flux of a propagating sinusoidal plane wave is proportional to the square of the time-derivative of the fundamental field. In electromagnetism, the Poynting flux is proportional to the square of the time-derivative of the vector potential.
In general relativity, the flux is therefore proportional to the square of the time-derivative of $`h(t)`$. The proportionality constant must be built only out of $`c`$, $`G`$, and pure numbers. To get the right units, it must be proportional to $`c^3/G`$; to get the pure number, a calculation in general relativity is required: $`1/32\pi `$ for a linearly polarized wave. (Polarization is described later.) This gives
$`F_{gw}`$ $`=`$ $`{\displaystyle \frac{1}{32\pi }}{\displaystyle \frac{c^3}{G}}\left({\displaystyle \frac{dh}{dt}}\right)^2`$
$`=`$ $`1.6\times 10^5\left({\displaystyle \frac{f}{100\mathrm{Hz}}}\right)^2\left({\displaystyle \frac{h}{10^{22}}}\right)^2\mathrm{W}\mathrm{m}^2,`$
for a wave with frequency $`f=\omega /2\pi `$. For comparison, reflected sunlight from Jupiter has a flux on Earth of $`2.3\times 10^7\mathrm{W}\mathrm{m}^2`$, almost 100 times smaller than that of a gravitational wave with an amplitude of $`10^{22}`$!
### Deficiencies of the quasi-Newtonian model
The calculation and equations in this section have been framed within a modified Newtonian model of gravity with a propagation speed of $`c`$, and one would expect some differences from general relativity. The most important difference is in the direction in which the tidal forces act. In the simple model, wave accelerations act in the $`z`$-direction, which was the direction of propagation of the wave. This is called a longitudinal wave.
In general relativity, gravitational waves are transverse waves: if the wave propagates in the $`z`$-direction then the tidal forces act only in the $`xy`$-plane. We will discuss later the exact form that their action in this plane takes. Remarkably, the rest of the formulas above are good approximations even in general relativity, provided $`ฯต`$ is calculated correctly, as described in the next section.
## The emission of gravitational waves
The previous section described the propagation of gravitational waves, their interaction with detectors, and the energy they carry. This section deals with the strength with which waves are emitted by astronomical bodies.
In Newtonian gravity there is a fundamental theorem, proved by Newton, that the gravitational field outside a spherical body is not only spherical, but it is the same as that of a point mass located at the origin of the body. It has the form given in Equation (1). In particular, the field is independent of the size of the body, as long as we consider only points outside it. This is true even if the star pulsates in a spherical manner.
This theorem is essentially the same in general relativity, and is known as Birkhoffโs Theorem. Outside a spherical body the field is the same as that of a black hole of the same mass as the body (the Schwarzschild metric), even if the body is pulsating spherically. But if the pulsation is nonspherical, then the outside field will change. In general relativity the changes generally propagate as a wave. So gravitational waves will be emitted by nonspherical motions.
In general the calculation of the emitted waves is extremely difficult, since the field equations of general relativity are a system of many coupled, nonlinear, partial differential equations. But in four circumstances the emission mechanisms are understood in some detail:
* Small-amplitude pulsations of relativistic stars and black holes. Normally gravitational radiation carries away energy and damps pulsations away, but in rotating stars the opposite may happen: the radiated loss of angular momentum may allow the star to spin down to an energetically more favored state, in which case the perturbation will grow, at least until nonlinear effects intervene. Discovered by S. Chandrasekhar and now called the Chandrasekhar-Friedman-Schutz (CFS) instability, it is thought to limit the rotation speed of young neutron stars (see below). Black holes also emit gravitational radiation when they are disturbed, e.g. by something falling into them, but they are not unstable: they always settle down into a steady state again.
* Radiation from โtestโ objects orbiting black holes. If the mass of the object is small enough then the total gravitational field may be treated as a linear perturbation of the exactly known field of a black hole (the Schwarzschild or, with rotation, the Kerr solution). These studies give insight into the general problem of gravitational radiation, and they also predict gravitational waveforms that might be observed by space-based observatories looking at compact stars falling into the giant black holes in the centers of galaxies. (See below and the article Gravitational Radiation Detection on Earth and in Space.)
* Weak gravitational fields and slow motion. Such weakly relativistic sources are studied in the post-Newtonian approximation, which includes higher-order corrections to Newtonian gravity from general relativity. This is analogous to the slow-motion multipole approximation that is so powerful in the study of electromagnetic radiation. Most realistic gravitational-wave sources can be studied to some approximation this way.
* Collisions of black holes and neutron stars. These events, which are expected to be observed by gravitational wave detectors (see below), must be modeled by solving the full set of Einstein equations on a powerful computer. Techniques to do this are advancing rapidly, and simulations of realistic mergers of stars and black holes from in-spiraling orbits can be expected to yield useful results in the first years of the 21st century.
### Quadrupole approximation
The post-Newtonian approximation has so far been the most powerful of these methods, and it yields the most insight into the emission mechanisms. Its fundamental result is the quadrupole formula, which gives the first approximation to the radiation emitted by a weakly relativistic system.
The quadrupole formula is analogous to the dipole formula of electromagnetism. In this language, monopole means spherical, which emits no radiation. This is also true in electromagnetism, where it is linked to conservation of charge. The โmonopole momentโ in electromagnetism is the total charge of a system, and since that does not change, there can be no spherical radiation.
Again in electromagnetism, the dipole moment is defined as the integral
$$d_i=\rho x_id^3x,$$
where $`\rho `$ is the charge density and $`x_i`$ is a Cartesian coordinate. If this integral is time-dependent, then the amplitude of the electromagnetic waves will be proportional to its first time-derivative $`dd_i/dt`$, and the radiated energy will be proportional (as we remarked earlier) to the square of the time derivative of this amplitude, i.e. to $`_i|d^2d_i/dt^2|^2`$.
In the post-Newtonian approximation to general relativity, the calculation goes remarkably similarly. The monopole moment is now the total mass-energy, which is the dominant source of the gravitational field for non-relativistic bodies, and which is constant as long as the radiation is weak. (Radiation will carry away energy, but in the post-Newtonian approximation that is a higher-order effect.) The dipole moment is given by the same equation as above, but with $`\rho `$ interpreted as the density of mass-energy.
However, here general relativity departs from electromagnetism. The time-derivative of the dipole moment is, since the mass-energy is conserved, just the integral of the velocity $`v_i`$:
$$\dot{d}_i=\rho v_id^3x.$$
(10)
But this is the total momentum in the system, and (to lowest order) this is constant. Therefore, there is no energy radiated due to dipole effects in general relativity. The gravitational field far from the source does contain a dipole piece if $`\dot{d}_i`$ is non-zero, but this is constant because it reflects the fact that the source has non-zero total momentum and is therefore moving through space.
To find genuine radiation in general relativity one must go one step beyond the dipole approximation to the quadrupole terms. These are also studied in electromagnetism, and the analogy with relativity again is close. The fundamental quantity is the spatial tensor (matrix) $`Q_{jk}`$, the second moment of the mass (or charge) distribution:
$$Q_{jk}=\rho x_jx_kd^3x.$$
(11)
A gravitational wave in general relativity is represented by a matrix $`h_{jk}`$ rather than a single scalar $`h`$, and its source (in the quadrupole approximation) is $`Q_{jk}`$.
As in electromagnetism, the amplitude of the radiation is proportional to the second time-derivative of $`Q_{jk}`$, and it falls off inversely with the distance $`r`$ from the source. A factor of $`G/c^4`$ is needed in order to get a dimensionless amplitude $`h`$, and a factor of 2 to be consistent with the definition in Equation (8). The result for $`h_{jk}`$ is:
$$h_{jk}=\frac{2G}{rc^4}\frac{d^2Q_{jk}}{dt^2}.$$
(12)
General relativity describes waves with a matrix because gravity is geometry, and the effects of gravity are represented by the stretching of space-time. This matrix contains that distortion information. Here is the information about the transverse action of the waves that the quasi-Newtonian model of the last section did not get right.
### Simple estimates
If the motion inside the source is highly non-spherical, then a typical component of $`d^2Q_{jk}/dt^2`$ will (from Equation (11)) have magnitude $`Mv_{N.S.}^2`$, where $`v_{N.S.}^2`$ is the non-spherical part of the squared velocity inside the source. So one way of approximating any component of Equation (12) is
$$h\frac{2GMv_{N.S.}^2}{rc^4}.$$
(13)
Comparing this with Equation (8) we see that the ratio $`ฯต`$ of the wave to the Newtonian potential is simply
$$ฯต\frac{v_{N.S.}^2}{c^2}.$$
By the virial theorem for self-gravitating bodies, this will not be larger than
$$ฯต<\varphi _{int}/c^2,$$
(14)
where $`\varphi _{int}`$ is the maximum value of the Newtonian gravitational potential inside the system. This provides a convenient bound in practice. It should not be taken to be more accurate than that.
For a neutron star source one has $`\varphi _{int}0.2c^2`$. If the star is in the Virgo cluster, then the upper limit on the amplitude of the radiation from such a source is $`5\times 10^{22}`$. This has been the goal of detector development for decades, to make detectors that can observe waves at or below an amplitude of $`10^{21}`$.
### Polarization of gravitational waves
The matrix nature of the wave amplitude comes from general relativity and has no Newtonian analog. In order to find the effect of the waves on the separation of two free particles (the idealized detector), one has to start with $`h_{jk}`$ as given by Equation (12) or by any other calculation, and then do three things:
1. Project the matrix $`h_{jk}`$ onto a plane perpendicular to the direction of travel of the wave. In the simple case considered above, where the wave was traveling in the $`z`$-direction, this means leaving the components $`\{h_{xx},h_{xy},h_{yy}\}`$ alone and setting the remaining components to zero. It is then a two-dimensional matrix in the transverse plane.
2. Remove the two-dimensional trace of the projected matrix. Call the resulting matrix $`h_{jk}^{TT}`$, where TT stands for Transverse-Traceless. In the example this means subtracting $`(h_{xx}+h_{yy})/2`$ from both $`h_{xx}`$ and $`h_{yy}`$. Then there are only two independent components left, $`h_{xy}^{TT}=h_{yx}^{TT}`$ and $`h_{xx}^{TT}=h_{yy}^{TT}`$.
3. To find the change in the separation of two particles that have an initial separation given by the vector $`\mathrm{}_k`$, let the matrix $`h_{jk}^{TT}`$ act on it:
$$\delta \mathrm{}_j=\underset{k}{}h_{jk}^{TT}\mathrm{}_k.$$
(15)
It is clear that any longitudinal component of the separation $`\mathrm{}_j`$ between the particles is unaffected by the wave (in the example, this is the $`z`$-separation), and that there are two degrees of freedom (the two independent components of $`h_{jk}^{TT}`$) to move particles in the plane perpendicular to the propagation direction. These two degrees of freedom are the two polarizations of the wave.
Fig. 1 shows the conventional definition of the two independent polarizations, from which any other can be made by superposition. What is shown is the effect of a wave on a ring of free particles in a plane transverse to the wave. The first line shows a wave with $`h_{xy}=0`$, conventionally called the โ+โ polarization. The bottom line shows a wave with $`h_{xx}=0`$, the โ$`\times `$โ polarization.
### Luminosity in gravitational waves
The energy carried by the gravitational wave must be proportional to the square of the time-derivative of the wave amplitude, so it will depend on the sum of the squares of the components $`d^3Q_{jk}/dt^3`$. The energy flux falls off as $`1/r^2`$, but when integrated over a sphere of radius $`r`$ to get the total luminosity, the dependence on $`r`$ goes away, as it should. The luminosity contains a factor $`G/c^5`$ on dimensional grounds, and a further factor of $`1/5`$ comes from a careful calculation in general relativity. The result is the gravitational wave luminosity in the quadrupole approximation:
$$L_{gw}=\frac{G}{5c^5}(\underset{j,k}{}\stackrel{\mathrm{}}{\text{Q}}_{jk}\stackrel{\mathrm{}}{\text{Q}}_{jk}\frac{1}{3}\stackrel{\mathrm{}}{\text{Q}}^2),$$
(16)
where $`Q`$ is the trace of the matrix $`Q_{jk}`$. Its squared third derivative must be subtracted in order to ensure that spherical motions do not radiate.
This equation will be used in the next section to estimate the back-reaction effect on a system that emits gravitational radiation.
## Emission estimates
Until observations of gravitational waves are successfully made, one can only make intelligent guesses about most of the sources that will be seen. There are many that could be strong enough to be seen by the early detectors: binary stars, supernova explosions, neutron stars, the early Universe. The estimates in this section are accurate only to within factors of order 1. The estimates correctly show how the important observables scale with the properties of the systems.
### Man-made gravitational waves
One source can be ruled out: man-made gravitational radiation. Imagine creating a wave generator with the following extreme properties. It consists of two masses of $`10^3`$ kg each (a small car) at opposite ends of a beam 10 m long. At its center the beam pivots about an axis. This centrifuge rotates 10 times per second. All the velocity is non-spherical, so $`v_{N.S.}^2`$ in Equation (13) is about $`10^5\mathrm{m}^2\mathrm{s}^2`$. The frequency of the waves will actually be 20 Hz, since the mass distribution of the system is periodic with a period of 0.05 s, only half the rotation period. The wavelength of the waves will therefore be $`1.5\times 10^7`$ M, about the diameter of the earth. In order to detect gravitational waves, not near-zone Newtonian gravity, the detector must be at least one wavelength from the source. Then the amplitude $`h`$ can be deduced from Equation (13): $`h5\times 10^{43}`$. This is far too small to contemplate detecting!
### Radiation from a spinning neutron star
Some likely gravitational wave sources behave like the centrifuge, only on a grander scale. Suppose a neutron star of radius $`R`$ spins with a frequency $`f`$ and has an irregularity, a bump of mass $`m`$ on its otherwise axially symmetric shape. Then the bump will emit gravitational radiation (again at frequency $`2f`$ because it spins about its center of mass, so it actually has mass excesses on two sides of the star), and the non-spherical velocity will be just $`v_{N.S.}=2\pi Rf`$. The radiation amplitude will be, from Equation (13),
$$h_{bump}2(2\pi Rf/c)^2Gm/rc^2,$$
(17)
and the luminosity, from Equation (16) (assuming roughly 4 comparable components of $`Q_{jk}`$ contribute to the sum),
$$L_{bump}(G/5c^5)(2\pi f)^6m^2R^4.$$
The radiated energy would presumably come from the rotational energy of the star. This would lead to a spindown of the star on a timescale
$$t_{spindown}=\frac{1}{2}mv^2/L_{bump}\frac{5}{4\pi }f^1\left(\frac{Gm}{Rc^2}\right)^1\left(\frac{v}{c}\right)^3.$$
It is felt that neutron star crusts are not strong enough to support asymmetries with a mass of more than about $`m10^5M_{}`$, and from this one can estimate the likelihood that the observed spindown timescales of Pulsars are due to gravitational radiation. In most cases, it seems that gravitational wave losses cannot be the main spindown mechanism.
But lower levels of radiation would still be observable by detectors under construction, and this may be coming from a number of stars. In particular, there is a class of neutron stars in X-Ray Binary Stars. They are accreting, and it is possible that accretion will create some kind of mass asymmetry or else lead to a rotational instability of the CFS type in the r-modes (see below). In either case, the stars could turn out be long-lived sources of gravitational waves.
### Radiation from a binary star system
Another โcentrifugeโ is a binary star system. Two stars of the same mass $`M`$ in a circular orbit of radius $`R`$ have $`v_{N.S.}^2=GM/4R`$. The gravitational-wave amplitude from Equation (13) can then be written
$$h_{binary}\frac{1}{2}\frac{GM}{rc^2}\frac{GM}{Rc^2}.$$
(18)
Compare this to the implications of putting Equation (14) into Equation (8).
The gravitational-wave luminosity of such a system is, by a calculation analogous to that for bumps on neutron stars,
$$L_{binary}\frac{1}{80}\frac{c^5}{G}\left(\frac{GM}{Rc^2}\right)^5.$$
In this equation there appears the important constant $`c^5/G=3.6\times 10^{52}`$ W, a number with the dimensions of luminosity built only from fundamental constants. By comparison, the luminosity of the Sun is only $`3.8\times 10^{26}`$ W. Close binaries can therefore radiate more energy in gravitational waves than in light.
The radiation of energy by the orbital motion causes the orbit to shrink. The shrinking will make any observed gravitational waves increase in frequency with time. This is called a chirp. The timescale for this is
$$t_{chirp}=Mv^2/L_{binary}\frac{20GM}{c^3}\left(\frac{GM}{Rc^2}\right)^4.$$
(19)
### The binary pulsar system โ verifying gravitational waves
This orbital shrinking has already been observed in the Hulse-Taylor Pulsar system, containing the radio pulsar PSR1913+16 and an unseen neutron star in a binary orbit. Discovered in 1974 by R Hulse and J Taylor, it has established that gravitational radiation is correctly described by general relativity. For their discovery, Hulse and Taylor received the 1993 Nobel Prize for Physics.
The key to the importance of this binary system is that all of the important parameters of the system can be measured before one takes account of the orbital shrinking. This is because a number of post-Newtonian effects on the arrival time of pulses at the Earth, such as the precession of the position of the periastron and the time-dependent gravitational redshift of the pulsar period as it approaches and recedes from its companion, are measured in this system. They fully determine the masses and separation of the stars and the inclination and eccentricity of their orbit. From these numbers, without any free parameters, it is possible to compute the shrinking timescale predicted by general relativity. The observed rate matches the predicted rate to within the observational errors of less than 1%.
The stars are in an eccentric orbit ($`e=0.615`$) and both have masses of $`1.4M_{}`$. The orbital semimajor axis is about $`7\times 10^8`$ m. Equation (19) assumes a circular orbit and gives a shrinking timescale of $`6\times 10^{10}`$ y. This is an overestimate, however, partly because it is in any case a rough approximation, and partly because the timescale is very sensitive to eccentricity. With the observed eccentricity, a careful calculation shows that the expected shrinking timescale is around $`4\times 10^8`$ y, consistent with observations.
### Chirping binaries
For a circular equal-mass binary, the orbital shrinking timescale and the frequency of the orbit determine both $`M`$ and $`R`$. If in addition a gravitational wave detector measures the wave amplitude $`h_{binary}`$, then the distance $`r`$ to the binary system can be determined.
Remarkably, this conclusion holds even for binaries with unequal masses. In such a case, the measurable mass is the chirp mass of the binary, defined as $`:=\mu ^{3/5}M_T^{2/5}`$, where $`\mu `$ is the reduced mass of the binary system and $`M_T`$ its total mass. Then the distance $`r`$ is still measurable from the chirp rate, frequency, and amplitude. In other words, a chirping binary is a standard candle in astronomy. Post-Newtonian corrections to the orbit, if observed in the waveform, can determine the individual masses of the stars and even their spins.
## Recognizing weak signals
For ground-based detectors, all expected signals have amplitudes that are close to or even below the instrumental noise level in the detector output. Such signals can nevertheless be detected with confidence if their waveform matches an expected waveform. The pattern recognition technique that will be used by detector scientists is called matched filtering.
Matched filtering works by multiplying the output of the detector by a function of time (called the template) that represents an expected waveform, and summing (integrating) the result. If there is a signal matching the waveform buried in the noise then the output of the filter will be higher than expected for pure noise.
A simple example of such a filter is the Fourier transform, which is a matched filter for a constant-frequency signal. The noise power in the data stream is spread out over the spectrum, while the power in the signal is concentrated in a single frequency. This makes the signal easier to recognize. The improvement of the signal-to-noise ratio for the amplitude of the signal is proportional to the square-root of the number of cycles of the wave contained in the data. This is well-known for the Fourier transform, and it is generally true for matched filtering.
Matched filtering can make big demands on computation, for several reasons. First, the arrival time of a short-duration signal is generally not known, so the template has to be multiplied into the data stream at each distinguishable arrival time. This is then a correlation of the template with the data stream. Normally this is done efficiently using Fast Fourier Transform methods.
Second, the expected signal usually depends on a number of unknown parameters. For example, the radiation from a binary system depends on the chirp mass $``$, and it might arrive with an arbitrary phase. Therefore, many related templates must be separately applied to the data to cover the whole family of signals.
Third, matched filtering enhances the signal only if the template stays in phase with the signal for the whole data set. If they go out of phase, the method begins to reduce the signal-to-noise ratio. For long-duration signals, such as for low-mass neutron-star coalescing binaries or continuous-wave signals from neutron stars (see below), this requires the analysis of large data sets, and often forces the introduction of additional parameters to allow for small effects that can make the signal drift out of phase with the template. It also means that the method works well only if there is a good prediction of the form of the signal.
Because the first signals will be weak, matched filtering will be used wherever possible. As a simple rule of thumb, the detectability of a signal depends on its effective amplitude $`h_{\text{eff}}`$, defined as
$$h_{\text{eff}}=h\sqrt{N_{cycles}},$$
(20)
where $`N_{cycles}`$ is the number of cycles in the waveform that are matched by the template.
For example, the effective amplitude of the radiation from a bump on a neutron star (Equation (17))will be $`h_{bump}\sqrt{2fT_{obs}}`$, where $`T_{obs}`$ is the observation time. In order to detect this radiation, detectors may need to observe for long periods, say 4 months, during which they accumulate billions of cycles of the waveform. During this time, the star may spin down by a detectable amount, and the motion of the Earth introduces large changes in the apparent frequency of the signal, so matched filtering needs to be done with care and precision.
Another example is a binary system followed to coalescence, i.e. where the chirp time in Equation (19) is less than the observing time. For neutron-star binaries observed by ground-based detectors this will always be the case (see the next section), so the effective amplitude is roughly
$$h_{chirp}h_{bin}\sqrt{f_{gw}t_{chirp}}\frac{GM}{rc^2}\left(\frac{GM}{Rc^2}\right)^{1/4},$$
(21)
where for $`f_{gw}`$ one must use twice the orbital frequency $`\sqrt{GM/R^3}/4\pi `$. This may seem a puzzling result, because it says that the effective amplitude of the signal gets smaller as the stars get closer. But this just means that the signal will be more detectable if it is picked up earlier, since Equation (21) assumes that the signal is followed right to coalescence. If one picks up the signal at earlier times, then there are more cycles of the waveform to filter for, and this naturally gives a better signal-to-noise ratio. This gives an advantage to detectors that can operate at lower frequencies. This has been an important consideration in the design of modern detectors. (See Gravitational Radiation Detection on Earth and in Space.)
In general, the sensitivity of detectors will be limited not just by detector technology, but also by the duration of the observation, the quality of the signal predictions, and even by the availability of computer processing power for the data analysis.
## Astronomical sources of gravitational waves
### Estimating the frequency
The signals for which the best waveform predictions are available have narrowly defined frequencies. In some cases the frequency is dominated by an existing motion, such as the spin of a pulsar. But in most cases the frequency will be related to the natural frequency for a self-gravitating body, defined as
$$f_0=\sqrt{G\overline{\rho }/4\pi },$$
(22)
where $`\overline{\rho }`$ is the mean density of mass-energy in the source. This is of the same order as the binary orbital frequency and the fundamental pulsation frequency of the body.
The frequency is determined by the size $`R`$ and mass $`M`$ of the source, taking $`\overline{\rho }=3M/4\pi R^3`$. For a neutron star of mass $`1.4M_{}`$ and radius 10 km, the natural frequency is $`f_0=1.9`$ kHz. For a black hole of mass $`10M_{}`$ and radius $`2GM/c^2=30`$ km, it is $`f_0=1`$ kHz. And for a large black hole of mass $`2.5\times 10^6M_{}`$, such as the one at the center of our Galaxy, this goes down in inverse proportion to the mass to $`f_0=4`$ mHz.
Fig. 2 shows the mass-radius diagram for likely sources of gravitational waves. Three lines of constant natural frequency are plotted: $`f_0=10^4`$ Hz, $`f_0=1`$ Hz, and $`f_0=10^4`$ Hz. These are interesting frequencies from the point of view of observing techniques: gravitational waves between 1 and $`10^4`$ Hz are accessible to ground-based detectors, while lower frequencies are observable only from space. (See the article Gravitational Radiaton Detection on Earth and in Space.) Also shown is the line marking the black-hole boundary. This has the equation $`R=2GM/c^2`$. There are no objects below this line. This line cuts through the ground-based frequency band in such a way as to restrict ground-based instruments to looking at stellar-mass objects. Nothing over a mass of about $`10^4M_{}`$ can radiate above 1 Hz.
A number of typical relativistic objects are placed in the diagram: a neutron star, a binary pair of neutron stars that spirals together as they orbit, some black holes. Two other interesting lines are drawn. The lower (dashed) line is the 1-year coalescence line, where the orbital shrinking timescale in Equation (19) is less than one year. The upper (solid) line is the 1-year chirp line: if a binary lies below this line then its orbit will shrink enough to make its orbital frequency increase by a measurable amount in one year. (In a one-year observation one can in principle measure changes in frequency of $`1\mathrm{yr}^1`$, or $`3\times 10^8`$ Hz.)
It is clear from the figure that any binary system that is observed from the ground will coalesce within an observing time of one year. Since pulsar statistics suggest that this happens less than once every $`10^5`$ years in our Galaxy, ground-based detectors must be able to register these events in a volume of space containing at least $`10^6`$ galaxies in order to have a hope of seeing occasional coalescences. When detectors reach this sensitivity (sometime in the first decade of the 21st century), then astronomers will be able to use the observed chirping binaries as standard candles to measure distance scales in the Universe.
### Radiation from neutron-star normal modes
In Fig. 2 there is a dot for the typical neutron-star. The corresponding frequency is the fundamental vibrational frequency of such an object. In fact, neutron stars have a rich spectrum of non-radial normal modes, which fall into several families: f-, g-, p-, w-, and r-modes have all been studied. If their gravitational wave emissions can be detected, then the details of their spectra would be a sensitive probe of their structure and of the Equation of State of Neutron Stars, in much the same way that Helioseismology probes the interior of the Sun. This is a challenge to ground-based detectors, which cannot yet make sensitive observations as high as 10 kHz.
### Radiation from gravitational collapse
The event that forms a neutron star is the gravitational collapse that produces in a supernova. It is difficult to predict the waveform or amplitude expected from this event, because we have no observational evidence about how nonspherical the collapse event might be in a typical supernova: the collapse is hidden deep within the star. So we can only guess. For example, a gravitational wave burst might be broad-band, centered on 1 kHz, or it might be a few cycles of radiation at a frequency anywhere between 100 Hz and 10 kHz, chirping up or down. The amplitude could be large, in which a good fraction of the energy released by the collapse is radiated in gravitational waves, or it could be negligibly small. It is indeed ironic that, although detecting supernovae was the initial goal of detector development when it started 4 decades ago, little more is known today about what to expect than scientists knew then.
### Radiation from r-modes
Hot neutron stars that rotate faster than about 100-200 Hz appear to be unstable to the emission of gravitational radiation through amplification of their r-modes by the CFS mechanism. In stars colder than about $`10^8`$ K, viscosity may be strong enough to damp out this instability. This instability may explain why only old, recycled, cold pulsars are seen at higher rotation rates. It also suggests that the formation of a rapidly rotating neutron star may be followed by a period of steady gravitational radiation as the star emits angular momentum and spins down to its stability limit. If as few as 10% of all the neutron stars formed since Star Formation began (at a redshift of perhaps 4) went through such a spindown, then they may have produced a detectable random background of gravitational radiation.
Interestingly, the r-modes are disturbances primarily of the fluid velocity; they have little density perturbation. Their name comes from their similarity to the Rossby waves of oceanography. The gravitational radiation they emit is not primarily mass-quadrupole (as in Equation (12)), but rather mass-current-quadrupole, the analog of magnetic quadrupole radiation in electromagnetism. This is the wave counterpart of what is called gravitomagnetism, which is responsible for the Lense-Thirring effect: an extra precession of a spinning gyroscope as it orbits a rotating body like the Earth caused by the spin-spin gravitational coupling of the gyroscope to the Earth.
## Black holes and gravitational waves
Black holes are regions of spacetime within which everything is trapped: light cannot escape, nor can anything else that moves slower than light. The boundary of this region is called the event horizon. This boundary is a dynamical surface. If any mass-energy falls into the hole, the area of the horizon increases. In addition, the horizon will generally wobble when this happens. These wobbles settle down quickly, emitting gravitational waves, and leaving a smooth (and slightly larger) horizon afterwards.
Undisturbed black holes are time-independent and smooth. In fact, according to general relativity the external gravitational field of such a black hole and the size and shape of its horizon are fully determined by only three numbers: the total mass, electric charge and angular momentum of the black hole. This black-hole uniqueness theorem is remarkable, considering how much variety there can be in the material that collapsed to form the black hole and that may have subsequently fallen in.
Observations of the gravitational waves emitted by a wobbling horizon or by a particle in orbit around a black hole have the potential to test the uniqueness theorem and thereby to verify the predictions of general relativity about the strongest possible gravitational fields.
Astronomers now recognize that there is an abundance of black holes in the universe. Observations of various kinds have located black holes in X-ray binary systems in the Galaxy and in the centers of galaxies.
These two classes of black holes have very different masses. Stellar black holes typically have masses of around $`10M_{}`$, and are thought to have been formed by the gravitational collapse of the center of a large, evolved Red Giant Star, perhaps in a supernova explosion. Massive black holes in galactic centers seem to have masses between $`10^6`$ and $`10^{10}M_{}`$, but their history and method of formation are not yet understood.
Both kinds of black hole can radiate gravitational waves. According to Fig. 2, stellar black hole radiation will be in the ground-based frequency range, while galactic holes are detectable only from space. The radiation from a black hole typically is strongly damped, lasting only a few cycles about the frequency, which for a spherical black hole is given by Equation (22) with $`R=2GM/c^2`$:
$$f_{BH}10\left(\frac{M}{M_{}}\right)^1\mathrm{Hz}.$$
### Stellar-mass black holes
Radiation from stellar black holes is expected mainly from coalescing binary systems, when one or both of the components is a black hole. Although such systems are thought to be rarer than systems of two neutron stars, the larger mass of the black hole makes the system visible from a greater distance. By measuring the chirp mass (as discussed above) observers will recognize that they have a black-hole system. It is very possible that the first observations of binaries by interferometers will be of black holes.
When a two-black-hole binary coalesces, there should be a burst of gravitational radiation that will depend in detail on the masses and spins of the objects. Numerical simulations of such events will be needed to interpret this signal, and possibly even to extract it from the instrumental noise of the detector. The research field of numerical relativity is making rapid progress, and it can be expected to produce informative simulations in the first few years of the 21st century, using the largest and fastest computers available at that time.
### Massive and supermassive black holes
Gravitational radiation is expected from supermassive black holes in two ways. In one scenario, two massive black holes spiral together in a much more powerful version of the coalescence we have just discussed. The frequency is much lower, but the amplitude is higher. Equation (21) implies that the effective signal amplitude is almost linear in the masses of the holes, so that a signal from two $`10^6M_{}`$ black holes will have an amplitude $`10^5`$ times bigger than the signal from two $`10M_{}`$ holes at the same distance. Even allowing for differences in technology, space-based detectors will be able to study such events with a very high signal-to-noise ratio no matter where in the universe they occur.
Observations of coalescing massive black-hole binaries will therefore provide strong tests of the validity of general relativity in the regime of strong gravitational fields, provided that numerical simulations can match the accuracy of the observations by that time.
The event rate for such coalescences is not easy to predict: it could be zero, but it may be large. It seems that the central core of most galaxies may contain a black hole of at least $`10^6M_{}`$. This is known to be true for our galaxy and for a number of others nearby. Supermassive black holes (up to a few times $`10^9M_{}`$) are believed to power Quasi-Stellar Objects and Active Galaxies. There is some evidence that the mass of the central black hole is proportional to the mass of the core of the host galaxy.
If black holes are formed with their galaxies, in a single spherical gravitational collapse event, and if nothing happens to them after that, then coalescences will never be seen. But it is believed that Galaxy Formation probably occurred through the merger of smaller units, sub-galaxies of masses upwards of $`10^6M_{}`$. If these units had their own black holes, then the mergers would have resulted in the coalescence of many of the holes on a timescale shorter than the present age of the universe. This would give an event rate of several per year. If the supermassive black holes were formed from smaller holes in a hierarchical merger scenario, then the event rate could be hundreds or thousands per year. It is likely that only space-based observations of gravitational waves will answer these questions.
A second scenario for the production of radiation by massive black holes is the swallowing of a stellar-mass black hole or a neutron star by the large hole. Massive black holes exist in the middle of dense star clusters. The tidal disruption of main-sequence or giant stars that stray too close to the hole is thought to provide the gas that powers the quasar phenomenon. These clusters will also contain a good number of neutron stars and stellar-mass black holes. They are too compact to be disrupted by the hole even if they fall directly into it.
Such captures therefore emit a gravitational wave signal that may be approximated by studying the motion of a โpoint massโ near a black hole. It will again emit a chirp of radiation, but in this case the orbit may be very eccentric. The details of the waveform encode information about the geometry of space-time near the hole. In particular, it may be possible to measure the mass and spin of the hole and thereby to test the uniqueness theorem for black holes. The event rate is not very dependent on the details of galaxy formation, and is probably high enough for many detections per year from a space-based detector.
## Gravitational waves from the Big Bang
Gravitational waves have traveled almost unimpeded through the universe since they were generated at times as early as $`10^{24}`$ s after the Big Bang. Observing them would provide important constraints on theories of Inflation and high-energy physics.
Inflation is an attractive scenario for the early universe because it makes the large-scale homogeneity of the universe easier to understand. It also provides a mechanism for producing initial density perturbations large enough to evolve into galaxies as the universe expands. These perturbations are accompanied by gravitational-field perturbations that travel through the universe, redshifting in the same way that photons do. Today these perturbations should form a random background of gravitational radiation.
The perturbations arise by parametric amplification of quantum fluctuations in the gravitational wave field that existed before inflation began. The huge expansion associated with inflation puts energy into these fluctuations, converting them into real gravitational waves with classical amplitudes.
If inflation did not occur, then the perturbations that led to galaxies must have arisen in some other way, and it is possible that this alternative mechanism also produced gravitational waves. One candidate is cosmic defects, including cosmic strings and cosmic texture. (See Topological Defects in Cosmology.) Although observations at present seem to rule cosmic defects out as a candidate for galaxy formation, cosmic strings may nevertheless have produced observable gravitational waves.
If inflation did not occur, there could also be a thermal background of gravitational waves at a temperature similar to that of the cosmological microwave background, but this radiation would have such a high frequency that it would not be detectable by any known or proposed technique.
The random background will be detectable as a noise in the detector that competes with instrumental noise. In a single detector, such as the first space-based detector, this noise must be larger than the instrumental noise to be detected, and one must have great confidence in the detector in order to claim that the observed noise is external. This is how the cosmic microwave background was originally discovered in a radio telescope.
If there are two detectors, then one can multiply the outputs of the two detectors together and sum (integrate). In this way, the random wave field in one detector acts like a matched filtering template, matching the random field in the other detector. This allows the detection of noise that is below the instrumental noise of the individual detectors. For this to work, the two detectors must be close enough together to experience the same random wave field. In practice, the sensitivity of this method falls off rapidly with separation if the detectors are more than a wavelength apart.
### Measure of the strength of random gravitational waves
When describing the strength of a random wave field, it is not appropriate to measure the amplitude of any single component. Rather, the r.m.s. amplitude of the field is the observable quantity. It is common to use an equivalent measure, the energy density $`\rho _{gw}(f)`$ in the radiation field as a function of frequency $`f`$. For a cosmological field, what is relevant is to normalize this energy density to the critical density $`\rho _c`$ required to close the universe. It is thus conventional to define
$$\mathrm{\Omega }_{gw}:=\frac{d\rho _{gw}/\rho _c}{d\mathrm{ln}f}.$$
(23)
This is roughly the fraction of the closure energy density in random gravitational waves between the frequency $`f`$ and $`2.718f`$.
Current and planned detectors may reach a sensitivity of $`\mathrm{\Omega }_{gw}10^9`$ at 1 mHz and $`10^{10}`$ at 40 Hz, but there is a possibility that backgrounds due to other sources (binary white dwarf systems and r-mode spindown, as discussed above) could obscure a cosmological background at these levels.
### Predicted spectrum of cosmological radiation
The simplest models of inflation suggest that the spectrum of the gravitational wave background should be flat, so that $`\mathrm{\Omega }_{gw}`$ is independent of frequency over a very large range of frequencies. In this case, the observed fluctuations in the cosmic microwave background radiation set a limit on gravitational radiation at ultra-low frequencies, and this constrains the energy density in the observable range (0.1 mHz to 10 kHz) to below about $`10^{13}`$ of closure. This will be too small to be seen by the current and planned detectors on the ground or in space.
But there is a great deal of room in these models for other spectra. The period before inflation may produce initial conditions for the phase of parametric amplification that give large amounts of radiation in the observable frequency range. One family of models based in superstring theory has a spectrum that rises at high frequencies. If a cosmological background from inflation or from cosmic defects can be observed, it will contain important clues to the nature of the theory that unifies gravitation with the rest of quantum physics.
## Conclusions
The first few years of the 21st century should see the first direct detections of gravitational radiation and the opening of the field of gravitational wave astronomy. Beyond that, over a period of a decade or more, one may expect observations to yield important and useful information about binary systems, stellar evolution, neutron stars, black holes, strong gravitational fields, and cosmology.
If gravitational wave astronomy follows the example of other fields, like X-Ray Astronomy and Radio Astronomy, then at some level of sensitivity it will begin to discover sources that were completely unexpected. Many scientists think the chance of this happening early is very good, since the processes that produce gravitational waves are so different from those that produce the electromagnetic radiation on which most present knowledge of the universe is based, and since more than 90% of the matter in the universe is dark and interacts with visible matter only through gravitation.
Present and planned detectors are known not to be ideal for some kinds of gravitational wave sources. Sensitive measurements of a cosmological background of radiation from the big bang may not be possible with these instruments if the spectrum follows the predictions of โstandardโ inflation theory. Most of the normal mode oscillations of neutron stars will be very hard to detect, because the radiation is weak and at a high frequency, but the science there is compelling: neutron-star seismology may be the only way to probe the interiors of neutron stars and understand these complex and fascinating objects. Detector technology will continually improve, and these sources provide important long-term goals for this field. There will clearly be much to do after the first observations are successfully made.
## Bibliography
Folkner, W M, ed., 1998 Laser Interferometer Space Antenna - AIP Conference Proceeding 456 (Woodbury, NY: American Institute of Physics)
Marck J-A, Lasota J-P, eds., 1996 Relativistic Gravitation and Gravitational Radiation (Cambridge and New York: Cambridge University Press)
Thorne, K S 1994 Black Holes and Time Warps: Einsteinโs Outrageous Legacy (New York: W. W. Norton and Co)
Wald, R M, ed., 1998 Black Holes and Relativistic Stars (Chicago: University of Chicago Press) |
warning/0003/math0003009.html | ar5iv | text | # When is the Fourier transform of an elementary function elementary?
## 1. Introduction
### 1.1. Motivations
Let $`F`$ be a local field, $`\psi :F^{}`$ a nontrivial unitary additive character, $`V`$ a finite dimensional vector space over $`F`$, and $`Q`$ a nondegenerate quadratic form on $`V`$. It is well known that the Fourier transform of the function $`\psi (Q)`$ has the form
(1.1)
$$\widehat{\psi (Q)}=ฯต(Q)\psi (Q^1)$$
where $`Q^1`$ is the inverse quadratic form on the dual space $`V^{}`$ (i.e., $`dQdQ^1=Id`$). As was shown in , Proposition 3, there exists an analog of (1.1) for some homogeneous rational functions on $`V`$ of homogeneity degree 2 which are not quadratic polynomials. More precisely, let $`E`$ be a cyclic cubic extension of $`F`$, and $`:F^{}^{}`$ a nontrivial cubic character which is trivial on the image of the norm map $`\mathrm{Nm}:E^{}F^{}`$. Let $`\varphi _{}`$ be the distribution on the vector space $`FE`$ such that
$$\varphi _{}(t,x)=(t)|t|^1\psi (\mathrm{Nm}(x)/t)$$
Then we have
(1.2)
$$\widehat{\varphi _{}}=ฯต\varphi _{},$$
where $`ฯต`$ is $`\pm 1`$. The proof of (1.2) given in is based on the analysis of the smallest special representation of the group $`D_4(F)`$ and uses global arguments (such as the existence of Eisenstein series). We were interested to see whether (1.2) could be proved by local methods and whether there exist interesting generalizations of (1.2). More precisely, let us say that a distribution $`g`$ on a vector space $`V`$ is โelementaryโ if it has the form $`C\psi (Q)\chi _1(P_1)\mathrm{}.\chi _k(P_k)`$ for some rational function $`Q`$ (called the phase function), polynomials $`P_1,\mathrm{},P_k`$ on $`V`$, and multiplicative characters $`\chi _1,\mathrm{},\chi _k`$ on $`F`$. We say that a distribution $`g`$ is โspecialโ if both $`g`$ and its Fourier transform $`\widehat{g}`$ are โelementaryโ. We were interested in finding which distributions are โspecialโ.
In this paper we present a number of โspecialโ distributions. Moreover, we show that the description of such distributions is almost independent of the local field $`F`$, and the Fourier transform $`\widehat{g}`$, up to a factor, is described algebraically in terms of $`g`$. Therefore this paper provides supporting evidence for the conjecture of the existence of an algebro-geometric integration theory proposed in .
### 1.2. Statement of the problem.
Let us formulate our main question more precisely. Keeping the notation of the previous section, set
$$G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}(x,\psi )=\psi (Q(x))\underset{j=1}{\overset{k}{}}\chi _j(P_j(x)).$$
The function $`G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}`$ is not always defined on the whole space $`V`$. However, if $`F`$ has characteristic zero (and conjecturedly, also in characteristic $`p`$), this function defines a distribution on $`V`$ which meromorphically depends on the characters $`\chi _j`$ (this follows from the resolution of singularities, or from the theory of D-modules in the archimedean case). In particular, for generic values of $`\chi _j`$ this function canonically extends to a distribution on $`V`$.
Question: For which $`Q,Q^{},P_i,P_j^{},\chi _i,\chi _j^{}`$ does one have
(1.3)
$$\widehat{G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}}=C_\psi G_{\chi _1^{},\mathrm{},\chi _l^{}}^{P_1^{},\mathrm{},P_l^{},Q^{}},\psi ,$$
as distributions on $`V^{}`$?
### 1.3. Results of the paper.
The main results of the paper are as follows.
In Section 2, using the formal stationary phase method, we give a necessary condition for (1.3), when $`Q`$ has nonzero Hessian. This condition says that $`Q^{}`$ is the Legendre transform of $`Q`$. We refer to this condition as โthe semiclassical conditionโ, commemorating the fact that identity (1.3) can be regarded as a โquantum mechanicalโ formula, from which this condition is deduced by using the โsemiclassicalโ (i.e., stationary phase) approximation.
In Section 3, we discuss properties and examples of phase functions satisfying the semiclassical condition. The most interesting examples we know come from prehomogeneous vector spaces. We classify phase functions such that both $`Q`$ and $`Q^{}`$ are of the form $`f(x)/t`$, where $`f`$ is a homogeneous cubic polynomial and $`t`$ an additional variable (Section 3). The classification says that in this case $`f`$ is a relative invariant of a regular prehomogeneous vector space of degree 3 (there are seven cases). The proof of this classification theorem is based on Zakโs classification theorem for Severi varieties.
In Section 4, we consider elementary functions in which the polynomials $`P_j`$ and the phase function $`Q`$ are monomial, and find a necessary and sufficient condition for the Fourier transform of such a function to be a function of the same type (in the โweakโ sense). This condition is an identity with $`\mathrm{\Gamma }`$-functions. We generalize this result to the case when $`Q,P_j`$ are monomials of norms of finite extensions of $`F`$, and, more generally, when they are monomials of relative invariants of prehomogeneous vector spaces over $`F`$.
In Section 5, we show that in the archimedean case, the condition of Section 4 for the existence of an integral identity can be reformulated in combinatorial terms (more precisely, in terms of so-called exact covering systems). We write down the explicit integral identities in the case when these combinatorial conditions are satisfied. We give the simplest nontrivial examples of integral identities, including the case of prehomogeneous vector spaces.
In Section 6, we generalize the results of Section 5 to non-archimedean fields $`F`$. Using the known formula for the Gamma function of a cyclic field extension, we obtain integral formulas of type (1.3). In particular, we give a new proof of formula (1.2).
These results have natural analogues in the case when $`F`$ is a finite field which will be described in a separate paper by D.K. and A.P. The role of distributions in this case is played by perverse sheaves and the Fourier transform is replaced by its geometric analogue defined by Deligne. Applying the trace of the Frobenius to identities with perverse sheaves, one obtains nontrivial elementary identities with exponential sums.
### 1.4. Acknowledgements.
We are grateful to D. Arinkin, and M. Kontsevich for useful discussions, and to B. Gross for calling our attention to Zakโs theorem. The work of P.E. and D.K. was supported by the NSF grant DMS-9700477. The work of A.P. was supported by the NSF grant DMS-9700458.
## 2. The semiclassical condition
### 2.1. Formulation of the semiclassical condition.
Recall the definition of the Legendre transform (see e.g., ). Let $`V`$ be a finite dimensional real vector space, $`v_0V`$, and $`Q`$ a smooth function on a neighborhood of $`v_0`$ such that $`\text{det}Q^{\prime \prime }(v_0)0`$. Let $`Q^{}(v_0)=p_0V^{}`$ (where $`Q^{},Q^{\prime \prime }`$ are the first and second differentials of $`Q`$). Then the Legendre transform of $`Q`$ is the smooth function $`L(Q)`$ defined in a neighborhood of $`p_0`$ by $`L(Q)(p)=pv_pQ(v_p)`$, where $`v_p`$ is the unique critical point of $`pvQ(v)`$ in a neighborhood of $`v_0`$.
This definition generalizes tautologically to the case when $`V`$ is a vector space over any field, and $`Q`$ is a regular function on the formal neighborhood of $`v_0V`$.
It is obvious that if $`Q`$ is an algebraic function then so is $`L(Q)`$.
Recall from Section 1.2 the definition of the function $`G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}(x,\psi )`$. For convenience we will always assume that the pole divisor of $`Q`$ is contained in the divisor $`P_1\mathrm{}P_k=0`$ (this does not cause a loss of generality).
Using the stationary phase method, we will prove the following theorem:
###### Theorem 2.1.
Let $`F`$ be $``$ or $``$. Suppose that (1.3) is satisfied, and $`Q`$ has nonzero Hessian. Then
(i) The rational map of algebraic varieties $`\underset{ยฏ}{V}\underset{ยฏ}{V}^{}`$ given by $`xQ^{}(x)`$ is a birational isomorphism.
(ii) $`Q^{}`$ is the Legendre transform of $`Q`$.
We will call this necessary condition of (1.3) the semiclassical condition.
Remark. We expect that a similar result holds over non-archimedean local fields.
Theorem 2.1 is proved in Section 2.3. In the next section, we explain the formal stationary phase method, which is necessary for the proof.
### 2.2. The formal stationary phase method.
The idea of the classical stationary phase method can be summarized as follows.
Let $`V`$ be a real finite dimensional vector space with a volume form. Let $`\varphi `$ be a function defined in an open set $`B`$ around $`0`$ in $`V`$ which has a nondegenerate critical point at $`0`$. Let $`f`$ be a smooth real-valued function whose support is a compact subset of $`B`$. Consider the integral
$$I(\mathrm{})=f(x)e^{i\varphi (x)/\mathrm{}}๐x,\mathrm{}>0.$$
###### Theorem 2.2.
(see and references therein) The function $`I(\mathrm{})`$ has the following asymptotic expansion as $`\mathrm{}0`$:
$$I(\mathrm{})C\mathrm{}^{dim(V)/2}|\text{det}(\varphi ^{\prime \prime }(0))|^{1/2}e^{i\varphi (0)/\mathrm{}}(f(0)+\underset{j=1}{\overset{\mathrm{}}{}}R_j(f,\varphi )(i\mathrm{})^j),$$
where $`R_j(f,\varphi )=\frac{\widehat{R}_j(f,\varphi )(0)}{\text{det}(\varphi ^{\prime \prime }(0))^{N_j}}`$, $`N_j_+`$, and $`\widehat{R}_j`$ are differential polynomials with rational coefficients.
Remark. The functions $`R_j`$ are complicated, but there is an algorithm of computing them which can be expressed in terms of Feynman diagrams.
Now let $`p`$ be a variable taking values in $`V^{}`$. If $`p`$ is small enough, the function $`\varphi (v)pv`$ has a unique critical point $`v_p`$ near zero, which is nondegenerate. Therefore, we have
$$f(x)e^{i(px\varphi (x))/\mathrm{}}๐x$$
$$C\mathrm{}^{dim(V)/2}|\text{det}(\varphi _p^{\prime \prime }(0))|^{1/2}e^{i\varphi _p(0)/\mathrm{}}(f_p(0)+\underset{j=1}{\overset{\mathrm{}}{}}R_j(f_p,\varphi _p)(i\mathrm{})^j),$$
where $`\varphi _p(v)=\varphi (v+v_p)p(v+v_p),f_p(v)=f(v+v_p)`$.
Note that $`\varphi _p(0)=L(\varphi )(p)`$, where $`L`$ is the Legendre transform.
Now we will generalize this to the formal setting. Let $`V`$ be a finite dimensional vector space over a field $`F`$ of characteristic zero. For any regular functions $`f,\varphi `$ on a formal neighborhood of zero in $`V`$, such that $`\varphi (0)=0,\varphi ^{}(0)=0`$, $`\text{det}(\varphi ^{\prime \prime }(0))0`$, define the regular function $`J_{f,\varphi }(h,p)`$ on the formal neighborhood of zero in $`V^{}((h))`$ by
$$J_{f,\varphi }(h,p)=\left(\frac{\text{det}(\varphi _p^{\prime \prime }(0))}{\text{det}(\varphi ^{\prime \prime }(0))}\right)^{1/2}e^{L(\varphi )(p)/h}(f_p(0)+\underset{j=1}{\overset{\mathrm{}}{}}R_j(f_p,\varphi _p)h^j)$$
(this is proportional to the right hand side of the stationary phase formula, with $`h=i\mathrm{}`$). We no longer claim that this series gives the asymptotic expansion of the integral $`f(x)e^{(\varphi (x)px)/h}๐x`$, because this integral is not defined. However, we can still claim that the series $`J`$ satisfies the same differential equations as the integral would satisfy if it existed. More precisely, we have the following lemma, which will be used to prove Theorem 2.1.
Let $`D`$ be a differential operator on $`V`$ with polynomial coefficients over $`F((h))`$, and let $`\widehat{D}`$ be the operator on $`V^{}`$ obtained from $`D`$ by the Fourier automorphism $`\frac{}{v}h^1v`$, $`ph\frac{}{p}`$.
Define the differential polynomial $`E_D(f,\varphi )`$ by $`D(fe^{\varphi /h})=E_D(f,\varphi )e^{\varphi /h}`$.
###### Lemma 2.3.
One has
$$\widehat{D}J_{f,\varphi }(h,p)=J_{E_D(f,\varphi ),\varphi }(h,p).$$
Proof. The statement is obvious from the stationary phase formula if $`F=`$ (by integration by parts), and $`f,\varphi `$ are expansions of smooth functions such that $`f`$ has compact support inside of the domain of $`\varphi `$. Since the statement is purely algebraic, it holds in general (because an arbitrary jet can be the jet of a function with compact support). โ
### 2.3. Proof of Theorem 2.1.
We will consider the case $`F=`$; the case of $``$ is similar.
Let $`Q,P_j,\chi _j`$, $`j=1,..,k`$ be as in Section 1.1. Let $`ZV_{}`$ be the locus of zeros of $`P_i`$. The function $`g(x):=G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}(x,\psi )`$ is smooth and nonvanishing on $`V_{}Z`$ and hence generates a 1-dimensional local system on $`V_{}Z`$. Let us extend this local system to an irreducible D-module on $`V_{}`$ and call this extension $`M_g`$. (Note that all our D-modules are algebraic D-modules).
Consider the D-module $`M(r)`$ generated by the distribution $`g_r:=(P_1\mathrm{}P_k)^rg`$ for sufficiently large $`r`$. Since $`M(r)`$ is holonomic, it has finite length, and so for sufficiently large $`r`$, the D-module $`M(r)`$ is independent of $`r`$. Let us denote this D-module by $`M_{\mathrm{}}`$.
It is easy to see that there is an exact sequence of D-modules
$$0KM_{\mathrm{}}M_g0,$$
where $`K`$ is supported on the divisor $`P_1\mathrm{}P_k=0`$. This implies that $`M_g`$ is isomorphic to $`D_V/I_r`$ (for large enough $`r`$), where $`D_V`$ is the algebra of differential operators on $`V`$, and $`I_r`$ is the left ideal of differential operators which annihilate $`g_r`$ formally (i.e., outside of $`Z`$).
###### Proposition 2.4.
For generic $`\psi `$, the rank of the Fourier transform of $`M_g`$ is at least the degree $`d`$ of the map $`Q^{}:\underset{ยฏ}{V}\underset{ยฏ}{V^{}}`$.
Proof. Let us change the ground field from $``$ to $`K=\overline{((\mathrm{}))}`$, and set formally $`\psi (x)=e^{ix/\mathrm{}}`$. It is enough to prove the claim for this particular $`\psi `$.
For this purpose, it is enough to produce, for a generic $`p_0V_{}^{}`$, a collection of $`d`$ linearly independent solutions of the differential equations $`\widehat{D}\varphi =0`$, $`DI_r`$, in the formal neighborhood of $`p_0`$.
To produce such solutions, we will use the formal stationary phase method. We will pick $`p_0`$ generically. Then the equation $`Q^{}(x)=p_0`$ has exactly $`d`$ distinct solutions $`x_1,\mathrm{},x_d`$, and $`Q^{\prime \prime }`$ is nondegenerate at all these points. Define the power series $`\varphi _i(x)=Q(x+x_i)Q(x_i)p_0x,f_i(x)=_j\chi _j(P_j)(P_1\mathrm{}P_k)^r(x+x_i)`$. Define
$$\eta _j(\mathrm{},p)=J_{f_j,\varphi _j}(i\mathrm{},p).$$
It follows from Lemma 2.3 that these series are indeed solutions of the equations $`\widehat{D}(p+p_0)\varphi (p)=0`$, $`DI_r`$.
It remains to show that the solutions $`\eta _i`$ are linearly independent. To do this, consider the power series $`L(\varphi _i)(p)`$, and look at their second degree terms, which are equal to $`Q^{\prime \prime }(x_i)^1(p,p)`$.
###### Lemma 2.5.
For generic $`p_0`$, the forms $`Q^{\prime \prime }(x_i)^1`$ are distinct.
Proof. The forms $`Q^{\prime \prime }(x_i)^1`$ are all the values at $`p=p_0`$ of the multivalued algebraic function $`((Q^{})^1)^{}(p)`$, which is the derivative of the function $`(Q^{})^1(p)`$ that has exactly $`d`$ branches by the definition.
But we claim that the derivative of any algebraic vector-function has at least (and hence exactly) as many branches as the function itself. Indeed, it is enough to check it for scalar functions $`f(z)`$ of one variable. But in the one variable case, one always has $`f(z,f^{})`$, since any monic algebraic equation of degree $`>1`$ satisfied by $`f`$ over $`(z,f^{})`$ can be differentiated to get a monic equation of lower degree. The lemma is proved. โ
Now let us make a change of variables $`\mathrm{}t^2\mathrm{}`$, $`ptp`$, and let $`t`$ tend to $`0`$. Then we have $`\eta _je^{iQ^{\prime \prime }(x_j)^1(p,p)/\mathrm{}}`$, which are linearly independent functions since $`Q^{\prime \prime }(x_j)^1`$ are distinct. This implies that $`\eta _i`$ are linearly independent. The proposition is proved. โ
Let $`G_1=G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}`$, $`G_2=G_{\chi _1^{},\mathrm{},\chi _l^{}}^{P_1^{},\mathrm{},P_l^{},Q^{}}`$.
###### Corollary 2.6.
Suppose that (1.3) holds. Then for generic $`\psi `$ the Fourier transform of the D-module $`M_{G_1}`$ is isomorphic to $`M_{G_2}`$.
Proof. The relation $`\widehat{G}_1=C_\psi G_2`$ implies that the Fourier transform of the D-module generated by $`G_1`$ is the D-module generated by $`G_2`$. Both of these D-modules are holonomic and have only one component of the Jordan-Holder series which has nonzero rank, namely, $`M_{G_1}`$ and $`M_{G_2}`$. But by Proposition 2.4, the rank of the Fourier transform of $`M_{G_1}`$ is positive. This implies the statement. โ
Now let us prove the theorem. We start with statement (i). By Proposition 2.4, for generic $`\psi `$ the rank of the Fourier transform $`M_{G_1}`$ is at least $`d`$. On the other hand, by Corollary 2.6 this Fourier transform is $`M_{G_2}`$, so its rank is $`1`$. So $`d=1`$, as desired.
Statement (ii) follows from the fact that $`G_2`$ satisfies (outside of $`Z`$) the differential equations $`\widehat{D}f=0`$, where $`DI_r`$, and hence (since the rank of the Fourier transform of $`M_{G_1}`$ is $`1`$), the formal expansion of $`G_2`$ must coincide with the formal series constructed using the stationary phase method. This implies that $`Q^{}=L(Q)`$, since it is clear that if $`J_{f,\varphi }(h,p)=e^{\xi (p)/h}(1+O(p,h))`$ then $`\xi =L(\varphi )`$ (this follows from theorem 2.2). The theorem is proved.
Remark. It is known that Theorem 2.2 has a non-archimedean analogue, which is even simpler than this theorem itself: in this case, the asymptotic expansion of the integral contains only the leading term and no higher terms. We expect that this result can be used to prove Theorem 2.1 in the non-archimedean case.
### 2.4. Integral identities in the weak sense and the semiclassical condition.
Let $`F`$ be archimedean, and $`V`$ a finite dimensional vector space over $`F`$. Let $`P`$ be a polynomial on $`V`$ and $`R`$ a polynomial on $`V^{}`$. Let $`N`$ a positive integer, and $`๐ฎ_N^{P,R}(V)`$ be the space of Schwartz functions on $`V`$ of the form $`|P|^{2N}f`$ (where $`f`$ is a Schwartz function), whose Fourier transform has the form $`|R|^{2N}g`$ (where $`g`$ is a Schwartz function). It is easy to construct examples of elements of this space: for instance, one can take the function $`|R|^{2N}()|P|^{2N^{}}(x)f`$, where $`N^{}>>N`$, and $`f`$ is any Schwartz function.
As we mentioned before, the function $`G_{\chi _1,\mathrm{},\chi _k}^{P_1,\mathrm{},P_k,Q}`$ defines a distribution only for generic values of the characters $`\chi _j`$. However, for any characters $`\chi _j`$ this function defines a linear functional on the space $`๐ฎ_N^{P_1\mathrm{}P_k,1}(V)`$ for large enough $`N`$.
Definition. We will say that the integral identity (1.3) holds in the weak sense if it holds on the space $`๐ฎ_N^{P_1^{}\mathrm{}P_l^{},P_1\mathrm{}P_k}(V^{})`$ for large enough $`N`$.
###### Theorem 2.7.
Theorem 2.1 remains valid if (1.3) holds only in the weak sense.
The proof is analogous to the proof of Theorem 2.1.
## 3. Rational functions satisfying the semiclassical condition.
In this section we would like to study systematically the question: which rational functions satisfy the semiclassical condition?
### 3.1. Properties of functions satisfying the semiclassical condition.
Let $`V`$ be a finite dimensional vector space over an algebraically closed field $`F`$. Denote by $`SC(V)`$ the set of rational functions $`Q`$ on $`V^{}`$ satisfying the semiclassical condition, i.e., such that $`Q^{}:VV^{}`$ is a birational isomorphism.
One can characterize elements of $`SC(V)`$ using the notion of Legendre transform, as follows.
###### Proposition 3.1.
A function $`QF(V)`$ belongs to $`SC(V)`$ if and only if the Legendre transform of $`Q`$ is rational.
Proof. The โonly ifโ part is obvious. To prove the โifโ part, let $`L(Q)=G`$ and let us differentiate the equation $`xQ^{}(x)Q(x)=G(Q^{}(x))`$. We get $`xQ^{\prime \prime }(x)=G^{}(Q^{}(x))Q^{\prime \prime }(x)`$. Since $`Q^{\prime \prime }`$ is generically nondegenerate, we get $`x=G^{}(Q^{}(x))`$. Thus, $`G^{}`$ is the inverse to $`Q^{}`$. โ
Nondegenerate quadratic forms are the simplest examples of elements of $`SC(V)`$. In the following sections we will construct other examples of elements of $`SC(V)`$.
### 3.2. The projective semiclassical condition.
Definition. A homogeneous rational function $`f`$ on $`V`$ is said to satisfy the projective semiclassical condition if the map $`xf^{}(x)`$ defines a birational isomorphism $`VV^{}`$.
Denote the set of functions satisfying the projective semiclassical condition by $`PSC(V)`$.
The relationship between $`SC(V)`$ and $`PSC(V)`$, which motivates the introduction of $`PSC(V)`$, is given by the following easy lemma.
###### Lemma 3.2.
Let $`g:VW`$ be a homogeneous of degree $`d`$ rational map of finite dimensional vector spaces, which defines a birational isomorphism $`\overline{g}:VW`$. Then $`g`$ itself is a birational isomorphism if and only if $`d=\pm 1`$. In particular, an element $`QPSC(V)`$ belongs to $`SC(V)`$ if and only if its homogeneity degree is $`0`$ or $`2`$.
Proof. It is enough to prove the first statement; the second statement is a special case of the first one for $`g=Q^{}`$.
If. The condition of the lemma implies that for a generic vector $`v`$ one has $`v=tR(g(v))`$, where $`R`$ is a rational function, and $`t`$ is a factor to be determined. Thus we have $`t^dg(R(g(v)))=g(v)`$, which allows one to determine $`t`$ rationally since $`d=\pm 1`$.
Only if. This part is clear, since a homogeneous birational isomorphism between vector spaces has to have homogeneity degree $`1`$ or $`1`$. โ
###### Corollary 3.3.
A homogeneous function $`fF(V)`$ belongs to $`PSC(V)`$ if and only if the map $`f^{}/f:VV^{}`$ is a birational isomorphism.
Proof. The โifโ part is clear. The โonly ifโ part follows when one applies Lemma 3.2 to $`g=f^{}/f`$. โ
###### Corollary 3.4.
(i) Any nonzero (in $`F`$) integer power of a function $`fPSC(V)`$ belongs to $`PSC(V)`$.
(ii) Let $`V,W`$ be finite dimensional $`F`$-vector spaces, and $`fPSC(V)`$, $`gPSC(W)`$. Then the exterior tensor product $`(fg)(v,w)=f(v)g(w)`$ on $`VW`$ belongs to $`PSC(VW)`$.
(iii) Any function of the form $`f_1^{n_1}\mathrm{}f_k^{n_k}`$, where $`f_iPSC(V_i)`$, and $`n_i`$ are nonzero in $`F`$, belongs to $`PSC(_iV_i)`$; it belongs to $`SC(_iV_i)`$ iff $`n_id_i=0`$ or $`2`$, where $`d_i`$ are the homogeneity degrees of $`f_i`$.
Proof. Statements (i) and (ii) are obvious from Corollary 3.3. Statement (iii) follows from (i),(ii), and Lemma 3.2. โ
Part (iii) of Corollary 3.4 allows one to obtain functions satisfying the semiclassical condition from functions satisfying the projective semiclassical condition. The simplest example of functions so obtained are monomial functions $`x_1^{n_1}\mathrm{}x_k^{n_k}`$, where $`n_i=0`$ or $`2`$.
So we will now study the projective semiclassical consition more systematically.
### 3.3. The multiplicative Legendre transform.
To think about elements of $`PSC(V)`$, it is useful to introduce the notion of the multiplicative Legendre transform.
Let $`f`$ be a homogeneous function, and $`det((f^{}/f)^{})`$ is not identically zero. In this case we can define a function $`f_{}`$ by $`f_{}(f^{}/f(x))=1/f(x)`$ (as the usual Legendre transform, it can be defined in an analytic as well as a formal setting). If $`f`$ is homogeneous of degree $`d`$ then so is $`f_{}`$.
Definition. We will call $`f_{}`$ the multiplicative Legendre transform of $`f`$.
Remark. Our terminology is motivated by the fact that $`f_{}`$ is the multiplicative Legendre transform of $`f`$ if and only if $`L(\mathrm{ln}f)=d+\mathrm{ln}f_{}`$ (over $``$).
It is obvious that the operation $`ff_{}`$ commutes with exterior tensor product. Also, $`(f^n)_{}=n^{nd}f_{}^n`$, where $`d`$ is the degree of $`f`$ (if $`n0`$ in $`F`$).
Example. $`(x_i^{n_i})_{}=n_i^{n_i}x_i^{n_i}`$.
###### Proposition 3.5.
(i) $`f_{}^{}/f_{}f^{}/f=Id`$.
(ii) $`f_{}=f`$.
Proof. The first identity is obtained by differentiating the definition of $`f_{}`$. The second one is obtained by applying $`f_{}`$ to both sides of (i). โ
One can characterize elements of $`PSC(V)`$ using the notion of the multiplicative Legendre transform, as follows.
###### Proposition 3.6.
A homogeneous rational function $`f`$ belongs to $`PSC(V)`$ if and only if $`f_{}`$ is rational.
Proof. The โonly ifโ part follows from Corollary 3.3. The โifโ part follows from Proposition 3.5 (i). โ
Remark. We see from the above that elements of $`PSC(V)`$ are multiplicative analogs of elements of $`SC(V)`$.
Example. Let us point out an easy method of creating new functions satisfying the projective semiclassical condition out of ones already known. It is straightforward to compute that if $`f_{}`$ is the multiplicative Legendre transform of $`f`$ on $`V`$ then the multiplicative Legendre transform of the function $`๐ฝ(x,y)=f^{}(x)y+f(x)`$ on $`V^2`$ is $`๐ฝ_{}(x_{},y_{})=(d1)^{1d}(f_{}^{}(y_{})x_{}f_{}(x_{}))^{d1}f_{}(y_{})^{2d}`$. This formula is valid also for $`d=1`$ if we agree that $`0^0=1`$.
Remark. Note that if $`๐ฝ`$ is a polynomial and $`d3`$ then $`๐ฝ_{}`$ is not a polynomial. Construction of polynomial elements $`fPSC(V)`$ such that $`f_{}`$ is also a polynomial is more tricky, and the only examples we know are described in the next section.
### 3.4. Construction of elements of $`PSC(V)`$ from prehomogeneous vector spaces.
Recall that a prehomogeneous vector space over $`F`$ is a triple $`(G,V,\chi )`$, where $`G`$ is an algebraic group over $`F`$, $`V`$ an algebraic representation of $`G`$, and $`\chi `$ a nontrivial algebraic character of $`G`$ such that
(i) $`V`$ has a Zariski dense $`G`$-orbit, and
(ii) there exists a nonzero polynomial $`f`$ on $`V`$ such that $`f(gv)=\chi (g)f(v)`$, $`gG`$ (it is obvious that if condition (i) holds, $`f`$ is unique up to a scalar).
Here we will assume that the group $`G`$ is reductive.
Remark. Prehomogeneous vector spaces were introduced in the 60-s by Sato (see ). Prehomogeneous vector spaces over $``$ with reductive $`G`$ and irreducible representation $`V`$ have been classified, see .
Let $`G_0=Ker(\chi )`$. Then $`G_0`$ is a codimension $`1`$ subgroup of $`G`$, and the function $`f`$ is invariant under $`G_0`$. The following proposition characterizes the ring of all $`G_0`$-invariants.
###### Proposition 3.7.
The ring of $`G_0`$-invariants of $`V`$ is $`F[f]`$.
###### Corollary 3.8.
$`G_0`$ has a dense orbit in $`V`$.
Proof. (of Proposition 3.7)
Let $`h`$ be any homogeneous polynomial which is an eigenfunction for $`G`$ and invariant under $`G_0`$. Since $`G/G_0=๐พ_m`$, there exist nonzero integers $`k,l`$ such that $`h^kf^l`$ is $`G`$-invariant. Since the $`G`$-action has a dense orbit, this implies that $`h^kf^l`$ is a constant, i.e., $`h,f`$ are powers of the same polynomial $`g`$, which is invariant under $`G_0`$ and is an eigenfunction of $`G`$. By the definition of $`f`$, this polynomial has to be $`f`$ (we canโt have $`g^k=f`$ for $`k>1`$ since then there would exist elements of $`G`$ that are not in $`G_0`$ but preserve $`f`$). The proposition is proved. โ
The function $`f`$ (which is uniquely determined up to scaling) is called the relative invariant of $`(G,V,\chi )`$.
From now till the end of the subsection we will assume that the characteristic of $`F`$ is zero.
Recall that a regular prehomogeneous vector space is such that $`det(f^{\prime \prime })`$ is not identically zero.
###### Proposition 3.9.
(see ) Let $`(G,V,\chi )`$ be a regular prehomogeneous vector space, and $`f`$ its relative invariant. Then $`fPSC(V)`$, and its multiplicative Legendre transform is a polynomial.
Proof. It is clear that $`(G,V^{},\chi ^1)`$ is a prehomogeneous vector space. Indeed, the existence of an open orbit, and the existence of a relative invariant of the same degree $`d`$ follows from the fact that the representation $`S^kV^{}`$ is completely reducible.
Let $`f_{}`$ be the relative invariant of $`(G,V^{},\chi ^1)`$. Consider the function $`f_{}f^{}`$ on $`V`$. This function is nonzero because of regularity, has degree $`d(d1)`$, and is $`G_0`$-invariant. Thus, this function is proportional to $`f^{d1}`$. So $`f_{}`$ is the multiplicative Legendre transform of $`f`$ up to scaling. The proposition is proved. โ
Examples. 1. $`G=GL(1)`$, $`V=F`$, $`\chi (a)=a^n`$, $`f(x)=x^n`$.
2. $`G=GL(1)^n`$, $`V=F^n`$, the action is $`(a_1,\mathrm{},a_n)(x_1,\mathrm{},x_n)=(a_1x_1,\mathrm{},a_nx_n)`$, $`\chi (a)=a_1\mathrm{}a_n`$, $`f=x_1\mathrm{}x_n`$.
3. $`G=GL(n)`$, $`V=S^2F^n`$, where $`F^n`$ is the vector representation, $`\chi =det^2`$, $`f=det`$.
4. $`G=GL(n)\times GL(n)`$, $`V=F^n(F^n)^{}`$, $`\chi =detdet^1`$, $`f=det`$.
5. $`G=GL(2n)`$, $`V=\mathrm{\Lambda }^2F^{2n}`$, $`\chi =det`$, $`f=Pf`$ (the Pfaffian).
6. $`G=E_6\times GL(1)`$, $`V`$ is the 27-dimensional irreducible representation (with $`GL(1)`$ acting by scalar multiplication), $`\chi `$ is $`zz^3`$, $`f`$ is the invariant cubic form.
7. $`G=O(N)\times GL(1)\times GL(1)`$, $`V=F^NF`$, where $`F^N`$ is the vector representation of $`O(N)`$, the action $`(g,a,b)(v,x)=(agv,bx)`$, $`\chi (g,a,b)=a^2b`$, $`f(v,x)=Q(v)x`$, where $`Q`$ is the invariant quadratic form.
To conclude this subsection, we would like to raise two questions.
Question 1. Is it true that any polynomial $`fPSC(V)`$ such that $`f_{}`$ is also a polynomial, is a relative invariant of a prehomogeneous vector space?
Question 2. Is it true that any polynomial $`fPSC(V)`$ such that $`f_{}`$ is also a polynomial, is rigid? In other words, is the set of equivalence classes of such polynomials finite for each degree?
Remark. Clearly, a positive answer to question 1 implies a positive answer to question 2, but question 2 could be more tractable.
For degree $`3`$, the answer to both questions is yes. This is proved in the next subsection.
### 3.5. Classification of cubic forms with cubic multiplicative Legendre transform.
In this section the field $`F`$ has characteristic zero. We prove the following theorem.
###### Theorem 3.10.
Let $`f`$ be a cubic form on a finite dimensional vector space $`V`$ such that its multiplicative Legendre transform is also a cubic form. Then $`f`$ is given by one of Examples 1-7 of the previous section.
The rest of the section is the proof of this theorem.
(Warning: It was pointed out by P. Sabatino and F. Viviani that this proof is incomplete. Namely, the argument with the Hessian at the end of the proof of Proposition 3.16 is not, by itself, sufficient to conclude that $`Z0`$ is smooth. However, a different proof of Theorem 3.10 has been given in .)
First of all, we may assume that $`f`$ (and $`f_{}`$) are irreducible (in which case $`dim(V)3`$). If any of them is reducible, it is a product of a linear and a quadratic form or of three linear forms, and it is easy to see that it is given by examples 1,2, or 7.
The functions $`f,f_{}`$ satisfy the equations
(3.1)
$$f_{}(f^{})=f^2,f(f_{}^{})=f_{}^2,f^{}f_{}^{}(x)=f_{}(x)x,f_{}^{}f^{}(x)=f(x)x.$$
Let $`XV`$ be the zero locus of $`f`$. Let $`ZX`$ be the zero locus of $`(f,f^{})`$, i.e., the singular locus of $`X`$. Define $`Z_{}X_{}V^{}`$ in a similar way.
###### Lemma 3.11.
(i) $`Z=(f^{})^1(0)`$, $`Z_{}=(f_{}^{})^1(0)`$.
(ii) $`X=(f^{})^1(Z_{})`$, $`X_{}=(f_{}^{})^1(Z)`$.
Proof. Part (i) follows from the fact that $`f^{}=0`$ implies $`f=0`$, and similarly for $`f_{}`$ (Euler equation). Part (ii) follows from equations (3.1). โ
Definition. A 3-dimensional subspace in $`V`$ is said to be a Cremona subspace if it contains three noncoplanar lines (through $`0`$) in $`Z`$ and is not entirely contained in $`X`$.
Cremona subspaces in $`V^{}`$ are defined similarly.
###### Lemma 3.12.
Let $`L`$ be a Cremona subspace in $`V`$. Then $`f^{}(L)`$ is a Cremona subspace in $`V^{}`$, and the map $`f^{}:Lf^{}(L)`$ is the Cremona map $`(x,y,z)(yz,zx,xy)`$ in some coordinates. In particular, the intersection of a Cremona subspace with $`Z`$ is the union of the three lines from the definition.
Proof. Let $`z_1,z_2,z_3`$ be noncoplanar lines in $`Z`$ that are contained in $`L`$, and let $`p_3=z_1z_2`$, $`p_2=z_1z_3`$, $`p_1=z_2z_3`$ be the planes spanned by pairs of lines. Consider the restriction of the map $`f^{}`$ to $`p_1`$. This map is given by a homogeneous quadratic form of two variables with values in $`V^{}`$, and it vanishes at lines $`z_2,z_3`$ by Lemma 3.11. Therefore, $`f^{}|_{p_1}`$ has the form $`q(v)w`$, where $`q`$ is a quadratic function, and $`wV`$. The same applies to $`p_2,p_3`$. Let $`z_1^{},z_2^{},z_3^{}`$ be the images of $`p_1,p_2,p_3`$. Then $`z_i^{}Z_{}`$ ($`p_iX`$ because $`X`$ is a cubic) and each $`z_i^{}`$ is a line or zero. Let $`L_{}`$ be a 3-subspace in $`V^{}`$ containing all $`z_i^{}`$. Let $`l`$ be a generic plane in $`L`$. Then $`f^{}(L)`$ has three intersection lines with $`L_{}`$ (images of intersections of $`l`$ with $`p_1,p_2,p_3`$). Since $`f^{}`$ is a quadratic map, we have $`f^{}(l)L_{}`$ and thus $`f^{}(L)L_{}`$.
Let us show that $`z_i^{}`$ are in fact nonzero (i.e. lines) and noncoplanar. If $`z_i^{}`$ lie in a 2-plane $`p`$, $`f_{}^{}f^{}|_L`$ is a map from a 3-space to a 2-plane. Therefore, $`L`$ must be contained in $`X`$, as it is obvious from (3.1) that $`f_{}^{}f^{}`$ is finite (4 to 1) outside of $`X`$. This contradicts the definition of $`L`$.
This shows that $`L_{}`$ is a Cremona subspace (it is clear that $`L_{}`$ is not entirely contained in $`X_{}`$), and it is obvious that if $`z_i,z_i^{}`$ are used as coordinate axes then $`f^{}`$ becomes the Cremona map. The lemma is proved. โ
###### Proposition 3.13.
A generic plane $`p`$ (through $`0`$) in $`V`$ is contained in a Cremona subspace.
Proof. Let $`a_1,a_2,a_3`$ be the three lines of intersection of $`p`$ with $`X`$. Let $`z_i^{}=f^{}(a_i)Z_{}`$. It is clear that $`z_i^{}`$ are noncoplanar, since otherwise $`pX`$, and we assumed that $`p`$ is generic. Thus the 3-space $`L_{}`$ spanned by $`z_i^{}`$ is a Cremona subspace. Let $`L=f_{}^{}(L_{})`$. Then by Lemma 3.12 $`L`$ is a Cremona subspace which contains $`a_i`$. โ
###### Proposition 3.14.
$`dim(Z)\frac{2}{3}dim(V)1`$, and $`Z`$ is not contained in any hyperplane.
Proof. Define a map $`g:Z\times Z\times Z\times F^3VV`$ by $`G(\zeta _1,\zeta _2,\zeta _3,b_1,b_2,b_3)=(\zeta _i,b_i\zeta _i)`$. This map is dominant, since by Proposition 3.13 two generic vectors are contained in a Cremona subspace. This implies the proposition. โ
###### Proposition 3.15.
$`Z`$ is irreducible.
Proof. It is enough to show that the map $`f_{}^{}:X_{}Z`$ is surjective, i.e., that for any $`\zeta 0`$, $`\zeta Z`$ there exists $`xX_{}`$ such that $`f_{}^{}(x)=\zeta `$. To do this, it is enough to show that any such $`\zeta `$ is contained in a Cremona subspace, since then $`\zeta `$ lies in the image of a special plane in the dual Cremona subspace in $`X_{}`$.
Suppose that $`\zeta _1,\zeta _2Z`$ are such that $`f^{\prime \prime }(\zeta )(\zeta \zeta _1,\zeta \zeta _2)0`$. Then $`f(\zeta +t(\zeta \zeta _1)+s(\zeta \zeta _2))`$ is of order $`ts`$ modulo cubic and higher terms at $`t,s=0`$, and thus the points $`0,\zeta ,\zeta _1,\zeta _2`$ are not in the same plane and span a Cremona subspace.
Thus, if the proposition was false, we would have $`f^{\prime \prime }(\zeta )(\zeta _1\zeta ,\zeta _2\zeta )=0`$ for all choices of $`\zeta _1,\zeta _2`$. But since $`Z`$ does not lie in any hyperplane, the possible vectors $`\zeta \zeta _1`$ span $`V`$, and similarly for $`\zeta \zeta _2`$. Thus, $`f(\zeta )=f^{}(\zeta )=f^{\prime \prime }(\zeta )=0`$. This implies that $`f(y+\zeta )=f(y)`$ for all $`yV`$, and hence $`f`$ is pulled back from $`V/<\zeta >`$. This is a contradiction, since we assumed that $`det(f^{\prime \prime })`$ is not identically zero. โ
###### Proposition 3.16.
$`dim(Z)=\frac{2}{3}dim(V)1`$ (in particular, $`d`$ is divisible by $`3`$), and $`Z0`$ is smooth.
Proof. Since $`Z`$ is irreducible, to prove the first statement it is sufficient to show that the map $`g`$ defined in the proof of Proposition 3.14 is generically finite. For this, it suffices to show that the Cremona subspace containing a generic plane is unique. Indeed, if $`p`$ is a generic plane, $`a_1,a_2,a_3`$ is as in the proof of Proposition 3.13, then $`a_i`$ are contained in any Cremona subspace containing $`p`$, so $`f^{}(a_i)=z_i^{}`$ are contained in the image of this subspace. But $`z_i^{}`$ are noncoplanar, so such a subspace is unique. The first statement is proved.
Let us now prove that $`Z0`$ is smooth. The dimension of the generic fiber of the map $`f_{}^{}:X_{}Z_{}Z0`$ is $`dim(X_{})dim(Z)=d/3`$, where $`d=dim(V)`$. Consider $`df_{}^{}:V^{}V`$. It is enough to show that the nullity of $`df_{}^{}=f_{}^{\prime \prime }`$ is $`d/3`$ at all points of $`X_{}Z_{}`$ (then it is exactly $`d/3`$ everywhere, and $`\pi `$ is a smooth fiber bundle).
Now we will need two simple lemmas.
Lemma 1. $`det(f_{}^{\prime \prime })=constf^{d/3}`$.
Proof. This follows from the fact that this determinant is nonzero outside of $`X`$ (where $`f_{}^{}`$ is a 2-1 covering), the irreducibility of $`f`$, and the fact that $`\text{deg}(\text{det}f_{}^{\prime \prime })=d`$.
Lemma 2. Suppose that $`A(t)`$ is a polynomial family of matrices such that $`detA(t)`$ vanishes exactly to the $`d`$-th order at $`t=0`$. Then the nullity of $`A(0)`$ is at most $`d`$.
Proof. We can assume that the kernel of $`A(0)`$ is the span of the first $`r`$ basis vectors, so the first $`r`$ columns of $`A`$ vanish at $`0`$. Thus, $`det(A)=O(t^r)`$, so $`rd`$.
Let us now apply Lemma 2 to $`f_{}^{\prime \prime }`$ restricted to a line transversal to $`X_{}`$ at a nonsingular point. Taking into account Lemma 1, we get that the nullity of $`f_{}^{\prime \prime }`$ at this point is $`d/3`$, which completes the proof of the proposition. <sup>2</sup><sup>2</sup>2Warning: Unfortunately, this argument is not sufficient to establish smoothness of $`Z0`$ (we thank P. Sabatino and F. Viviani for pointing this out). However, the proposition is valid, as is Theorem 3.10, as shown in , Corollary 4 by a different method.
Now we will finish the proof of the theorem. The above proposition shows that the projectivization $`(Z)`$ of $`Z`$ is a smooth projective variety of dimension $`\frac{2}{3}(dim(V)2)`$ in $`(V)`$, which does not lie in a hyperplane. It has the following property: the line connecting any two points of $`Z`$ is entirely contained in $`X`$ and thus never goes through a generic point of $`(V)`$. Varieties with these properties are called Severi varieties.
Now comes the central part of the proof, which is the use of the following classification theorem of Severi varieties, due to F. Zak.
###### Theorem 3.17.
Let $`Y`$ is a smooth, closed subvariety of the complex projective space $`P^{d1}`$, which does not lie in a hyperplane. Suppose that
(i) any line connecting two points of $`Y`$ belongs to a certain hypersurface, and
(ii) $`\text{dim}(Y)=\frac{2}{3}(d3)`$.
Then $`Y`$ is projectively equivalent to the singularity locus of the equation $`f=0`$ on $`(V)`$, where $`(V,f)`$ is one of the following four prehomogeneous vector spaces with relative invariant:
1. $`V`$ is the space of 3 by 3 symmetric matrices, $`f=\text{det}`$.
2. $`V`$ is the space of 3 by 3 matrices, $`f=\text{det}`$.
3. $`V`$ is the space of skew-symmetric 6 by 6 matrices, $`f=\text{Pf}`$.
4. $`V`$ is a 27-dimensional irreducible representation of $`E_6`$, $`f`$ is the invariant cubic form.
This theorem shows that in our situation, $`(V,f)`$ is given by one of the examples 3-6 in the previous section.
The variety $`X`$ is reconstructed from $`Z`$ as the set of points on lines connecting two points of $`Z`$. The theorem is proved. โ
###### Corollary 3.18.
Let $`Q(x,t)=f(x)/t`$, where $`f`$ is a homogeneous cubic polynomial on some finite dimensional complex vector space $`W`$, and $`t`$ is an additional variable. Then the following conditions are equivalent:
(i) $`QSC(W)`$, and $`Q^{}=\stackrel{~}{f}(x_{})/t_{}`$, where $`\stackrel{~}{f}`$ is a cubic polynomial on $`W^{}`$.
(ii) The polynomial $`f`$ is as in Examples 1-7.
Proof. It is easy to check directly that $`L(f(x)/t)=\stackrel{~}{f}(x_{})/t_{}`$ iff $`\stackrel{~}{f}=f_{}`$. Thus the statement follows from Theorem 3.10. โ
Remarks. 1. In examples 3-5, the map $`g`$ of Proposition 3.14 has a classical linear algebraic interpretation. Example 3 corresponds to simultaneous diagonalization of two quadratic forms. Example 4 corresponds to simultaneous diagonalization of two hermitian forms. Example 5 corresponds to simultaneous reduction of two skewsymmetric forms to the canonical form (sum of 2-dimensional forms).
2. Examples 2-7 correspond to the semisimple Jordan algebras of degree 3 . It is possible that one can check directly (i.e., without using Theorem 3.10) that in the assumptions of Theorem 3.10, if $`dim(V)>1`$ then $`V`$ admits a structure of a separable (hence semisimple) Jordan algebra of degree 3 such that $`f`$ is its determinant polynomial. This would allow to give another proof of Theorem 3.10 which would use Albertโs theorem on the classification of simple Jordan algebras, rather than Zakโs classification theorem.
## 4. Integral identities with monomials
### 4.1. Fourier transform.
In the next four sections we will recall some basic facts about analysis over local fields. The basic reference for these facts is the book .
Let $`F`$ be a local field. We fix a nontrivial additive character $`\psi `$. For any finite dimensional vector space $`V`$ over $`F`$, let $`๐ฎ(V)`$ be the space of (complex-valued) Schwartz functions on $`V`$.
For any Haar measure $`dx`$ on $`V`$, one defines the Fourier transform $`๐ฎ(V)๐ฎ(V^{})`$ by
(4.1)
$$\widehat{f}(y)=_F\psi (yx)f(x)๐x.$$
Any Haar measure on $`V`$ defines a positive inner product on $`๐ฎ(V)`$. Let us say that Haar measures $`dx`$ on $`V`$ and $`dx^{}`$ on $`V^{}`$ are compatible if the Fourier transform is a unitary operator with respect to this inner product. It is easy to see that this condition is symmetric, and that if it is satisfied then one has the inversion formula $`\widehat{\widehat{f}}(x)=f(x)`$.
If $`V`$ is identified with $`V^{}`$, one can choose a unique Haar measure which is compatible with itself. For example, this is so if $`V=F`$ or $`F^n`$. From now on, we will use the notation $`dx`$ for this special measure.
The measure $`dx`$ on $`F`$ depends on $`\psi `$. For example, in the archimedean case, if $`\psi (x)=e^{i\text{Re}(ax)}`$ then $`dx`$ is $`(|a|/2\pi )^{dim_{}F/2}`$ times the Lebesgue measure. Below, we use the character $`\psi (x)=e^{i\text{Re}(x)}`$ for $`F=,`$, and a character of norm $`1`$ for the non-archimedean case. This completely determines $`dx`$.
Let $`๐(V)`$ be the space of distributions on $`V`$. If $`V`$ carries a Haar measure $`dx`$, then along with the Fourier transform of Schwartz functions, one can define the Fourier transform of distributions $`๐(V^{})๐(V)`$ (by duality). It will also be denoted by $`f\widehat{f}`$. In the case of compatible measures, it satisfies the inversion formula.
### 4.2. Multiplicative characters.
Let $`F^{}`$ be the multiplicative group of $`F`$, $`U(F^{})`$ the group of continuous unitary characters of $`F^{}`$, and $`X(F^{})`$ the group of continuous characters of $`F^{}`$ into $`^{}`$. If $`F=`$, then $`U(F^{})=\times /2`$ and $`X(F^{})=\times /2`$. If $`F=`$, then $`U(F^{})=\times `$, and $`X(F^{})=\times `$. In the non-archimedean case, $`U(F^{})`$ is $`/\times D`$, where $`D`$ is a discrete countable group, and $`X(F^{})=/\times D`$.
Let $`d_mx`$ be the multiplicative Haar measure on $`F^{}`$, normalized so that $`d_mx/dx=1`$ at $`x=1`$. Let $`\nu _F(x)=(d_mx/dx)^1`$ be the norm of $`x`$ (in the archimedean case, it equals $`|x|^{\text{dim}_{}F}`$).
For $`\lambda X(F^{})`$, denote by $`\mathrm{Re}\lambda `$ the real number defined by $`|\lambda (x)|=\nu _F(x)^{\mathrm{Re}\lambda }`$. It is obvious that $`X(F^{})=U(F^{})\times `$, via $`\chi (\frac{\chi }{|\chi |},\mathrm{Re}\chi )`$. We think of the first coordinate as the imaginary part of $`\chi `$ and of the second as the real part of $`\chi `$.
Let us say that $`\lambda X(F^{})`$ is a singular character if $`\lambda (x)=\nu _F(x)^1`$ (in the non-archimedean case), $`\lambda (x)=\nu _F(x)^1x^n`$, $`n_0`$ (for $`F=`$), and $`\lambda (x)=\nu _F(x)^1x^n\overline{x}^m`$, $`n,m_0`$ (for $`F=`$). It is well known that $`\lambda (x)`$ is a holomorphic family of distributions on $`F`$ depending on $`\lambda X(F^{})`$, with simple poles at singular characters.
### 4.3. Gamma functions.
Now define the Gamma function $`\mathrm{\Gamma }^F(\lambda )`$ of a local field $`F`$ to be the meromorphic function on $`X(F^{})`$ given by
(4.2)
$$\widehat{\lambda \nu _F^1}(x)=\mathrm{\Gamma }^F(\lambda )\lambda ^1(x),$$
(whenever both $`\lambda \nu _F^1`$ and $`\lambda ^1`$ define distributions on $`F`$).
From the inversion formula for the Fourier transform one gets the functional equation
(4.3)
$$\mathrm{\Gamma }^F(\lambda )\mathrm{\Gamma }^F(\nu _F\lambda ^1)=\lambda (1)$$
Let us give the expressions for the Gamma functions of $``$ and $``$.
###### Lemma 4.1.
For $`s`$, let $`\lambda _{s,n}(x)=|x|^s(x/|x|)^n`$, $`xF`$. Let $`\mathrm{\Gamma }_n^F(s)=\mathrm{\Gamma }^F(\lambda _{s,n})`$. Then
(4.4)
$$\mathrm{\Gamma }_n^{}(s)=(2\pi )^{1/2}2i^n\mathrm{\Gamma }(s)\mathrm{cos}(\pi (sn)/2),n/2,$$
and
(4.5)
$$\mathrm{\Gamma }_n^{}(s)=(2\pi )^12^si^n\mathrm{\Gamma }(\frac{s+n}{2})\mathrm{\Gamma }(\frac{sn}{2})\mathrm{sin}(\pi (sn)/2),n.$$
This lemma is well known and is proved in a straightforward way.
One can also easily compute the Gamma function of a power of $`\nu _F`$ in the non-archimedean case. If $`q`$ is the order of the residue field of $`F`$ then
(4.6)
$$\mathrm{\Gamma }^F(\nu _F^s)=\frac{1q^{s1}}{1q^s}.$$
From these formulas it is clear that the Gamma function has simple poles at the characters $`\lambda \nu _F`$, where $`\lambda `$ is singular. It is clear from the definition that it is holomorphic everywhere else. (For an exact expression of the Gamma function in the non-archimedean case, see ).
### 4.4. Integral representation of the additive character.
Let $`du`$ be the Haar measure on $`U(F^{})`$ for which the Mellin transform $`L^2(F^{},d_mx)L^2(U(F^{}),du)`$ is a unitary operator.
We would like to consider a distribution on $`U(F^{})`$ of the form
$$\varphi _{U(F^{})}\mathrm{\Gamma }^F(u)\varphi (u)๐u.$$
Since $`\mathrm{\Gamma }^F(u)`$ has a pole at the trivial character, it is necessary to choose a regularization of this integral. Our choice, here and throughout, will be the following: to avoid the pole, the contour of integration in the connected component of the identity in $`X(F^{})`$ should be indented in the direction of positive values of $`\text{Re}(u)`$. The following lemma shows that this choice coincides with the choice of , where $`\mathrm{\Gamma }^F(u)`$ is defined as the Mellin transform of $`\psi `$.
###### Lemma 4.2.
The distribution $`\psi (x)`$ on $`F`$ has the following integral representation:
$$\psi (x)=_{U(F^{})}\mathrm{\Gamma }^F(u)u^1(x)๐u.$$
Remark. This integral is divergent for any concrete value of $`x`$ but is absolutely convergent on any test function from the Schwartz space.
Proof. It is easy to see from the definition of the Gamma function that the Fourier transform on the group $`F^{}`$ of the distribution $`\psi (x)`$ (i.e., the Mellin transform) is equal to $`\mathrm{\Gamma }^F`$ outside of $`0`$. Therefore the statement of the lemma on test functions which vanish at 0 is obtained by applying the inversion formula for the Fourier transform. It suffices now to check this identity on one test function which does not vanish at 0. In the non-archimedean case, this is easy to do for the characteristic function of the integers, and in the archimedean case one can do it for the function $`e^{|x|^2/2}`$. โ
### 4.5. The generalized Gamma function.
For a positive integer $`d`$, define a meromorphic function $`\mathrm{\Gamma }_{d,a}^F`$ on $`X(F^{})`$ by the formula
$$\mathrm{\Gamma }_{d,a}^F(\lambda )=d^1\underset{\mu :\mu ^d=\lambda }{}\mathrm{\Gamma }^F(\mu )\mu ^1(a).$$
We have $`\mathrm{\Gamma }_{1,a}^F(\lambda )=\mathrm{\Gamma }^F(\lambda )\lambda ^1(a)`$, so the function $`\mathrm{\Gamma }_{d,a}^F`$ is a generalization of the Gamma function. We will call $`\mathrm{\Gamma }_{d,a}^F`$ the generalized Gamma function.
Let us compute the generalized Gamma function in the archimedean case.
If $`F=`$ then
(4.7)
$$\mathrm{\Gamma }_{d,a}^F(\lambda _{s,nd})=\mathrm{\Gamma }^F(\lambda _{s/d,n})\lambda _{s/d,n}(a),$$
If $`F=`$ and $`d=2k+1`$ then
(4.8)
$$\mathrm{\Gamma }_{d,a}^F(\lambda _{s,n})=\frac{1}{d}\mathrm{\Gamma }^F(\lambda _{s/d,n})\lambda _{s/d,n}(a).$$
If $`F=`$, $`\psi (x)=e^{ix}`$, $`a>0`$, and $`d=2k`$ then
(4.9)
$$\mathrm{\Gamma }_{d,\pm a}^F(\lambda _{s,0})=(2\pi )^{1/2}k^1e^{\pm \pi is/4k}\mathrm{\Gamma }(s/2k)a^{s/2k}.$$
### 4.6. The distribution $`G_{\lambda _1,\mathrm{},\lambda _k,a}^{n_1,\mathrm{},n_k}`$ on $`F^k`$ and its integral representation.
Let $`n_1,\mathrm{},n_k`$ be integers, and $`\lambda _1,\mathrm{},\lambda _kX(F^{})`$. Consider the function on $`(F^{})^k`$ defined by
(4.10)
$$G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}(x_1,\mathrm{},x_k)=\psi (a\underset{i=1}{\overset{k}{}}x_i^{n_i})\lambda _1(x_1)\mathrm{}\lambda _k(x_k).$$
(Here $`aF^{}`$ is a parameter).
$`G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}`$ is a distribution on $`F^k`$ which is holomorphic in $`\lambda _i`$ if $`\mathrm{Re}\lambda _i>1`$.
###### Lemma 4.3.
For $`\lambda _i`$ with real parts $`>1`$ one has
$$G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}(x_1,\mathrm{},x_k)=_{U(F^{})}\mathrm{\Gamma }^F(u)u^1(a)\lambda _1u^{n_1}(x_1)\mathrm{}\lambda _ku^{n_k}(x_k)๐u.$$
(in the sense of distributions).
Proof. Follows directly from Lemma 4.2. โ
It will be convenient for us to understand the function $`G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,\mathrm{},n_k,a}`$ as a distribution in the weak sense. Namely, for polynomials $`P,R`$ define the space $`๐ฎ_N^{P,R}(V)`$ as in section 2 for the archimedean case, and as the space of Schwartz functions vanishing on the variety $`P=0`$ whose Fourier transform vanishes on $`R=0`$, in the non-archimedean case (so in the non-archimedean case it is independent of $`N`$). As in Section 2, for any $`\lambda _i`$ the function $`G_{\lambda _1,..,\lambda _k}^{n_1,\mathrm{},n_k,a}`$ defines a linear functional on the space $`๐ฎ_N^{P_1\mathrm{}P_k}(F^k)`$ for a large enough $`N`$. We call such a functional a distribution in the weak sense.
The following lemma is a straghtforward generalization of the previous lemma.
###### Lemma 4.4.
For any $`\lambda _i`$ one has
$$G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}(x_1,\mathrm{},x_k)=_{U(F^{})}\mathrm{\Gamma }^F(u)u^1(a)\lambda _1u^{n_1}(x_1)\mathrm{}\lambda _ku^{n_k}(x_k)๐u.$$
(as distributions in the weak sense).
### 4.7. The Fourier transform of the distribution $`G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}`$.
###### Lemma 4.5.
One has
(4.11) $`\widehat{G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}}(p_1,\mathrm{},p_k)=`$
$`{\displaystyle _{U(F^{})}}\mathrm{\Gamma }^F(u)u^1(a){\displaystyle \underset{i=1}{\overset{k}{}}}\mathrm{\Gamma }^F(\lambda _iu^{n_i}\nu _F)\mathrm{}\lambda _i^1u^{n_i}\nu _F^1(p_i)du.`$
Proof. This follows from the previous lemma. โ
### 4.8. Identities with monomials.
###### Theorem 4.6.
Let $`n_1,\mathrm{}n_k`$, $`m_1,\mathrm{},m_k`$ be nonzero integers. Let $`d=gcd(n_1,\mathrm{},n_k)`$. Then the identity
(4.12)
$$\widehat{G_{\lambda _1,\mathrm{}\lambda _k}^{n_1,\mathrm{},n_k,a}}=CG_{\eta _1,\mathrm{},\eta _k}^{m_1,\mathrm{},m_k,b},$$
between distributions on $`F^k`$ in the weak sense, is satisfied if and only if
(4.13)
$$\eta _i\lambda _i\nu _F=\gamma ^{n_i}.$$
where $`\gamma X(F^{})`$ is a character, and one of the following two conditions holds:
1. $`m_i=2`$, $`m_i=n_i`$, and
(4.14)
$$\mathrm{\Gamma }_{d,a}^F(u^d)\underset{i=1}{\overset{k}{}}\mathrm{\Gamma }^F(u^{n_i}\lambda _i\nu _F)=C\mathrm{\Gamma }_{d,b}^F(u^d\gamma ^d);$$
2. $`m_i=0`$, $`m_i=n_i`$, and
(4.15)
$$\mathrm{\Gamma }_{d,a}^F(u^d)\underset{i=1}{\overset{k}{}}\mathrm{\Gamma }^F(u^{n_i}\lambda _i\nu _F)=C\mathrm{\Gamma }_{d,b}^F(u^d\gamma ^d).$$
Proof. Using the previous two lemmas, we get
(4.16) $`{\displaystyle _{U(F^{})}}\mathrm{\Gamma }^F(u)u^1(a){\displaystyle \underset{i=1}{\overset{k}{}}}\mathrm{\Gamma }^F(\lambda _iu^{n_i}\nu _F)\lambda _i^1u^{n_i}\nu _F^1(p_i)du=`$
$`C{\displaystyle _{U(F^{})}}\mathrm{\Gamma }^F(v)v^1(b)\eta _1v^{m_1}(p_1)\mathrm{}\eta _kv^{m_k}(p_k)๐v.`$
We see that this identity can hold only if the vectors $`(m_1,\mathrm{},m_k)`$ and $`(n_1,\mathrm{},n_k)`$ are proportional. Let $`\alpha =m_i/n_i`$ for all $`i`$ (the proportionality coefficient). It is not difficult to see by asymptotic analysis of the above formula for $`u,v=\nu _F^s`$ for large $`s`$ (which is essentially equivalent to the โsemiclassical analysisโ of Section 2) that $`1m_i=\alpha `$. Therefore, $`1n_i=\alpha ^1`$, so both $`\alpha `$ and $`\alpha ^1`$ are integers and hence $`\alpha =\pm 1`$. Thus, $`m_i=\pm n_i`$.
So we should consider two cases.
Case 1. $`m_i=2`$, $`m_i=n_i`$. In this case, replacing $`v`$ with $`v^1`$, we obtain
(4.17) $`{\displaystyle _{U(F^{})}}\mathrm{\Gamma }^F(u)u^1(a){\displaystyle \underset{i=1}{\overset{k}{}}}\mathrm{\Gamma }^F(\lambda _iu^{n_i}\nu _F)\lambda _i^1u^{n_i}\nu _F^1(p_i)du=`$
$`C{\displaystyle _{U(F^{})}}\mathrm{\Gamma }^F(v^1)v(b)\eta _1v^{n_1}(p_1)\mathrm{}\eta _kv^{n_k}(p_k)๐v.`$
Let us replace $`u`$ with $`u^{1/d}`$. This leads to summation over all roots of degree $`d`$, and therefore the Gamma functions $`\mathrm{\Gamma }^F(u)`$, $`\mathrm{\Gamma }^F(v)`$ are replaced by the generalized Gamma functions:
(4.18) $`{\displaystyle _{U(F^{})}}\mathrm{\Gamma }_{d,a}^F(u){\displaystyle \underset{i=1}{\overset{k}{}}}\mathrm{\Gamma }^F(\lambda _iu^{n_i/d}\nu _F)\lambda _i^1u^{n_i/d}\nu _F^1(p_i)du=`$
$`C{\displaystyle _{U(F^{})}}\mathrm{\Gamma }_{d,b}^F(v^1)\eta _1v^{n_1/d}(p_1)\mathrm{}\eta _kv^{n_k/d}(p_k)๐v.`$
It is clear that the integrals on the two sides of this equation can coincide if and only if the contour of integration in the first integral can be shifted to obtain the second integral. In particular, there must exist a character $`\gamma `$ such that $`\eta _i\lambda _i\nu _F=\gamma ^{n_i}`$. In this case, replacing $`v`$ with $`u\gamma ^d`$ (i.e., shifting the contour of integration), we get
(4.19) $`{\displaystyle _{U(F^{})}}\mathrm{\Gamma }_{d,a}^F(u){\displaystyle \underset{i=1}{\overset{k}{}}}\mathrm{\Gamma }^F(\lambda _iu^{n_i/d}\nu _F)\lambda _i^1u^{n_i/d}\nu _F^1(p_i)du=`$
$`C{\displaystyle _{U(F^{})}}\mathrm{\Gamma }_{d,b}^F(u^1\gamma ^d)\lambda _1u^{n_1/d}\nu _F^1(p_1)\mathrm{}\lambda _ku^{n_k/d}\nu _F^1(p_k)๐u.`$
Remark. We can shift the contour of integration without worrying about residues, since our integral identities are understood in the weak sense, while residual contributions are distributions supported on the coordinate hyperplanes, which by definition do not affect identities in the weak sense.
The latter condition is equivalent to
(4.20)
$$\mathrm{\Gamma }_{d,a}^F(u)\mathrm{\Gamma }^F(\lambda _1u^{n_1/d}\nu _F)\mathrm{}\mathrm{\Gamma }^F(\lambda _ku^{n_k/d}\nu _F)=C\mathrm{\Gamma }_{d,b}^F(u^1\gamma ^d).$$
Since $`\mathrm{\Gamma }_{d,a}`$, by definition, vanishes away from the $`d`$-th powers, we can replace in this condition the variable $`u`$ by $`u^d`$ without changing the condition. This yields the equation (4.15) in the theorem.
Case 2. $`m_i=0`$, $`m_i=n_i`$. In this case, we see similarly to case 1 that (4.12) is equivalent to the combination of the two conditions from part 2 of the theorem.
The theorem is proved. โ
### 4.9. Gamma functions of prehomogeneous vector spaces.
The construction of the previous section can in fact be further generalized to a more general setting of prehomogeneous vector spaces.
For simplicity let $`F`$ have characteristic zero. Let $`\underset{ยฏ}{V}`$ be a prehomogeneous vector space of dimension $`M`$ over $`F`$ for a reductive group $`\underset{ยฏ}{G}`$. We assume that $`V=\underset{ยฏ}{V}(F)`$ has a dense orbit under the action of the group of points $`G=\underset{ยฏ}{G}(F)`$, and that the same is true for the dual prehomogeneous vector space $`V^{}`$.
Let $`f`$ be a relative invariant of $`V`$ generating its invariant ring. Suppose it has degree $`D`$. Let $`V^{}`$ be the dual space and $`f_{}`$ the relative invariant of $`V^{}`$, which is the multiplicative Legendre transform of $`f`$. Assume for simplicity that the stabilizer $`G_0`$ of $`f`$ in $`G`$ acts with finitely many orbits on the varieties $`f=0`$, $`f_{}=0`$ in $`V,V^{}`$.
For any multiplicative character $`\lambda `$ of $`F^{}`$, consider the function $`\lambda (f(x))`$ on $`V`$. Since $`F`$ has characteristic zero, it is known that this function defines a distribution on $`V`$ which meromorphically depends on $`\lambda `$.
It is easy to see that for generic $`\lambda `$ the Fourier transform $`\widehat{\lambda \nu _F^{\frac{M}{D}}(f)}(x)`$ is proportional to $`\lambda ^1(f_{}(x))`$. (For generic $`\lambda `$, this is the unique, up to scaling, distribution of the correct homogeneity degree, due to our assumption about the finiteness of the number of orbits). This allows one to define the Gamma function $`\mathrm{\Gamma }^V(\lambda )`$ of $`V`$ to be the meromorphic function on $`X(F^{})`$ given by
(4.21)
$$\widehat{\lambda \nu _F^{\frac{M}{D}}(f)}(x)=\mathrm{\Gamma }^V(\lambda )\lambda ^1(f_{}(x)),$$
(whenever both sides define distributions on $`F`$).
From the inversion formula for Fourier transform one gets the functional equation
(4.22)
$$\mathrm{\Gamma }^V(\lambda )\mathrm{\Gamma }^V(\nu _F^{M/D}\lambda ^1)=\lambda (1)^D.$$
In fact, it was shown by Sato and Shintani that if $`F`$ is archimedean, then these results are valid without the assumption of the finiteness of the number of orbits.
### 4.10. Identities with monomials on prehomogeneous vector spaces.
Let $`V_1,\mathrm{},V_k`$ be prehomogeneous vector spaces over $`F`$ as in the previous section, of dimensions $`M_i`$, and let $`f_i`$ be their relative invariants, of degrees $`D_i`$. Let $`V=V_i`$.
By a monomial on $`V`$ we will mean a rational function on $`V`$ of the form $`f_1(x_1)^{n_1}\mathrm{}f_k(x_k)^{n_k}`$, where $`n_i`$ are integers, and $`x_iV_i`$. Define the distribution on $`V`$ by the formula
(4.23)
$$๐ข_{\lambda _1,\mathrm{},\lambda _k}^{n_1,..,n_k,a}(x_1,\mathrm{},x_k)=\psi (a\underset{i=1}{\overset{k}{}}f_i(x_i)^{n_i})\lambda _1(f_1(x_1))\mathrm{}\lambda _k(f_k(x_k)).$$
Here $`aF^{}`$ is a parameter, and $`\lambda _iX(F^{})`$. Similar to the case of the usual monomials, this is a distribution for large real parts of $`\lambda _i`$ which meromorphically extends to generic $`\lambda _i`$, and a distribution in the weak sense for all $`\lambda _i`$.
For this class of distributions, we have the following generalization of Theorem 4.6.
###### Theorem 4.7.
Let $`n_1,\mathrm{}n_k`$, $`m_1,\mathrm{},m_k`$ be nonzero integers. Let $`d=gcd(n_1,\mathrm{},n_k)`$. Then the identity
(4.24)
$$\widehat{๐ข_{\lambda _1,\mathrm{}\lambda _k}^{n_1,\mathrm{},n_k,a}}=C๐ข_{\eta _1,\mathrm{},\eta _k}^{m_1,\mathrm{},m_k,b},$$
between distributions $`V^{}=V_i^{}`$ in the weak sense is satisfied if and only if
(4.25)
$$\eta _i\lambda _i\nu _F^{M_i/D_i}=\gamma ^{n_i},$$
where $`\gamma X(F^{})`$ is a character, and one of the following two conditions holds:
1. $`m_iD_i=2`$, $`m_i=n_i`$,
(4.26)
$$\mathrm{\Gamma }_{d,a}^F(u^d)\underset{i=1}{\overset{k}{}}\mathrm{\Gamma }^{V_i}(u^{n_i}\lambda _i\nu _F^{M_i/D_i})=C\mathrm{\Gamma }_{d,b}^F(u^d\gamma ^d);$$
2. $`m_iD_i=0`$, $`m_i=n_i`$, and
(4.27)
$$\mathrm{\Gamma }_{d,a}^F(u^d)\underset{i=1}{\overset{k}{}}\mathrm{\Gamma }^{V_i}(u^{n_i}\lambda _i\nu _F^{M_i/D_i})=C\mathrm{\Gamma }_{d,b}^F(u^d\gamma ^d).$$
Remark. Here the distribution $`๐ข_{\eta _1,\mathrm{},\eta _k}^{m_1,\mathrm{},m_k,b}`$ is defined using the relative invariants $`f_i^{}`$ of $`V_i^{}`$, which are multiplicative Legendre transforms of $`f_i`$.
Proof. The proof of this theorem is analogous to the proof for usual monomials. โ
An important special case of this theorem describes integral identities with monomials of norms of field extensions, which includes the identity from cited in Section 1.1. This special case is defined by setting $`V_i=F_i`$ (field extensions of $`F`$ of degrees $`d_i`$), and $`f_i=\mathrm{Nm}:F_iF`$ to be the norm maps. In this case, $`V_i`$ is identified with $`V_i^{}`$ using the trace functional.
### 4.11. Poles of distributions $`G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,\mathrm{},n_k,a}`$
We would like to find out when integral identities with monomials hold not only in the weak but also in the strong sense (i.e., on all test functions). For this purpose we should first of all find out where (in terms of $`\lambda _i`$) the distribution $`G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,\mathrm{},n_k,a}`$ could have poles. This is our goal in this section. For brevity we will denote this distribution just by $`G`$.
###### Proposition 4.8.
The divisor of poles of $`G`$ is a subset of the set of points where one of the following equations is satisfied:
(i) $`\lambda _i`$ is a singular character for some $`i`$ such that $`n_i0`$;
(ii) $`(\lambda _jr_j^1)^{n_l}=(\lambda _lr_l^1)^{n_j}`$ for some $`(j,l)`$ such that $`n_j>0>n_l`$, and singular characters $`r_j,r_l`$.
(The notion of a singular character was introduced in Section 4.2).
Proof. It is clear that we can assume that all the exponents $`n_i`$ are nonzero.
Suppose that a point $`(\lambda _1,\mathrm{},\lambda _k)X(F^{})^k`$ is a generic point of the divisor of poles of $`G`$. Then the leading coefficient $`G_{top}`$ of the Laurent expansion of $`G`$ at that point is a distribution on the coordinate cross. Its support is a closed subset $`S`$ of the coordinate cross which is invariant under the scaling group $`T=\{(t_1,\mathrm{},t_k)(F^{})^k:t_i^{n_i}=1\}`$.
Let $`H_I=\{x_i=0,iI\}`$, $`I\{1,\mathrm{},k\}`$. Assume that $`\lambda _i`$ is not a singular character whenever $`n_i>0`$. Then $`S_{I:|I|=k2}H_I`$.
Let $`y`$ be a generic point of $`S`$. In this case the $`T`$-orbit of $`y`$ spans a certain subspace $`H_I`$, $`|I|k2`$. In a small neighborhood of $`y`$, the sets $`S`$ and $`H_I`$ coincide.
Let $`T_I`$ be the group of elements of $`T`$ in which $`t_i=1`$, $`iI`$ (so $`T_I`$ acts trivially on $`H_I`$). On the one hand, for $`tT_I`$ we have $`G_{top}(tx)=_{iI}\lambda _i(t_i)G_{top}(x)`$. On the other hand, since in the neighborhood of $`y`$ the distribution $`G_{top}`$ is supported on $`H_I`$, there exist singular characters $`r_j`$, $`jI`$ such that $`G_{top}(tx)=_{iI}r_i(t_i)G_{top}(x)`$. This implies that whenever $`_{iI}t_i^{n_i}=1`$, we have $`_{iI}(\lambda _ir_i^1)(t_i)=1`$, for suitable singular characters $`r_i`$. In particular, for any distinct $`i,jI`$, one has $`(\lambda _ir_i^1)^{n_j}=(\lambda _jr_j^1)^{n_i}`$.
The last equation can define a component of the pole divisor only if $`n_j,n_l`$ have opposite signs, since otherwise there are solutions with $`\mathrm{Re}(\lambda _j)>0`$, $`\mathrm{Re}(\lambda _l)>0`$, where $`G`$ is holomorphic. The proposition is proved. โ
### 4.12. Identities with monomials in the weak and strong sense
Sometimes one can deduce from an integral identity in the weak sense that it actually holds in the strong sense (i.e., on all test functions). Let us do it in the case of archimedean fields, using the theory of D-modules.
Let $`F`$ be archimedean. Let us say that a nonsingular multiplicative character $`\lambda `$ is strongly regular if it generates (as a distribution) an irreducible D-module on the line. For $`F=`$ a character is strongly regular iff it is not of the form $`x^n/|x|`$ for an integer $`n`$, or $`x^n`$ for $`n<0`$. For $`F=`$, a character is strongly regular iff it is not of the form $`x^n\overline{x}^m`$, where $`n,m`$ are integers, and at least one of them is negative.
###### Theorem 4.9.
Suppose that for some k-tuples of characters $`(\lambda _1,\mathrm{},\lambda _k)`$, $`(\eta _1,\mathrm{},\eta _k)`$ identity (4.12) holds in the weak sense, and
(i) $`\lambda _i`$, $`\eta _i`$ are strongly regular whenever $`n_i0`$;
(ii) $`(\lambda _jr_j^1)^{n_l}(\lambda _lr_l^1)^{n_j}`$ for any $`(j,l)`$ such that $`n_j>0>n_l`$, and any singular characters $`r_j,r_l`$, and the same is true about $`\eta _i`$.
In this case both sides of (4.12) are well defined as distributions (on all test functions), and (4.12) is satisfied as an identity between distributions (in the โstrongโ sense).
Proof. According to Proposition 4.8, the distributions $`G_1=G_{\lambda _1,\mathrm{},\lambda _k}^{n_1,\mathrm{},n_k,a}`$, and $`G_2=G_{\eta _1,\mathrm{},\eta _k}^{m_1,\mathrm{},m_k,b}`$ are well defined. The fact that (4.12) holds in the weak sense means that $`\widehat{G}_1G_2=\mathrm{\Delta }`$, where $`\mathrm{\Delta }`$ is the sum of a distribution supported on the coordinate cross and a distribution whose Fourier transform is supported on the cross. Let $`M_i`$ be the D-modules generated by $`G_i`$.
Lemma. $`M_1`$ and $`M_2`$ are irreducible.
Proof. Let us prove that $`M_1`$ is irreducible. The irreducibility of $`M_2`$ is shown in a similar way.
Let $`I\{1,\mathrm{},k\}`$. We claim that if the numbers $`n_i,iI`$, have the same signs, then the D-module $`M_1`$ is irreducible on the formal polydisk around a generic point $`y`$ of the subspace $`H_I`$ (where $`H_I`$ was defined in the proof of Proposition 4.8). Indeed, if all $`n_i,iI`$ are positive, this restriction is isomorphic to the exterior tensor product of $`<\lambda _i>,iI`$ with $`k|I|`$ copies of $`<1>`$ (the structure sheaf of the 1-dimensional formal disk), so it is irreducible by assumption (i). If all $`n_i`$ are negative, the restriction is generated by $`_{jI}\lambda _j(x_j)e^{i\text{Re}(c(x)_{jI}x_j^{n_j})}`$, where $`c`$ depends only of $`x_j,jI`$, and is nonzero at $`y`$. This D-module is obviously irreducible for any $`\lambda _j`$.
Thus, we see that all Jordan-Holder components of $`M_1`$ whose support is contained in the cross, are supported on the union of subspaces $`H_{I_{j,l}}`$ for $`n_j>0>n_l`$, where $`I_{j,l}`$ is the complement of $`\{j,l\}`$. Now, using scaling arguments as in the proof of Proposition 4.8, it is easy to deduce from condition (ii) for $`j,l`$ that all these components are zero. This means that $`M_1`$ is irreducible. The lemma is proved. โ
Now let us prove the theorem. By the irreducibility of $`M_1`$, $`\mathrm{\Delta }`$ is supported on the cross (as well as its Fourier transform), since the restriction of $`M_1`$ to the set of its smooth points has to be an irreducible local system.
Lemma. The distribution $`\mathrm{\Delta }`$ has support of codimension 2 or more.
Proof. If the support of $`\mathrm{\Delta }`$ contains a point $`(x_1,\mathrm{},x_k)`$ with $`x_i=0`$, $`x_j0`$ for $`ji`$ then let us apply to $`\widehat{G}_1`$ the algebra of differential operators in one variable $`x_i`$, with coefficients in polynomials of other variables. It suffices to check that the space of obtained distributions contains a nonzero distribution supported on the cross (this would contradict the irreducibility of $`M_2`$).
For this, it suffices to check that for any distribution on the line of the form $`h=\chi (t)\psi (ct^m)+\mathrm{\Delta }`$, where $`m0`$, $`\chi `$ is strongly regular if $`m>0`$, and $`\mathrm{\Delta }`$ is supported at zero, $`\mathrm{\Delta }0`$, one can find a polynomial differential operator $`D`$ (algebraically depending on $`c`$) such that $`Dh`$ is nonzero but is supported at zero. This is straightforward. โ
Finally, since $`\text{Supp}(\mathrm{\Delta })`$ has codimension 2 or more, it is not difficult to see using homogeneity arguments that $`\mathrm{\Delta }=0`$.
The theorem is proved.โ
Remark. The same method can be used for identities with norms and prehomogeneous vector spaces; however, we do not give here the details of such arguments.
## 5. Identities over $``$ and $``$.
As we saw in the previous section, the Gamma functions for $``$ and $``$ are given by simple explicit formulas via the Euler Gamma function. One can also compute the Gamma functions of prehomogeneous vector spaces over these fields, using the theory of Bernstein polynomials. This allows one to express the Gamma function identities of the previous section in much more explicit (combinatorial) terms, thus giving a completely elementary criterion of the existence of integral identities. This is what we do in this section.
For the sake of brevity, we do a systematic analysis only in the case of ordinary monomials with relatively prime exponents, restricting ourselves to a number of examples in other cases.
### 5.1. The relation of the parameters $`a`$ and $`b`$.
###### Lemma 5.1.
Suppose that identity (4.26) is satisfied in the weak sense. Then if $`m_i=2`$, one has
(5.1)
$$ab=\underset{i}{}n_i^{n_iD_i},$$
and if $`m_i=0`$, one has
(5.2)
$$ab^1=\underset{i}{}n_i^{n_iD_i}$$
Proof. This follows from Theorem 2.1 or from analyzing the asymptotics of the Gamma function identity for large $`s`$. โ
### 5.2. Multiplication formulas for $`\mathrm{\Gamma }`$-functions.
We recall the classical multiplication law for $`\mathrm{\Gamma }`$-function:
(5.3)
$$\mathrm{\Gamma }(Nz)=N^{Nz\frac{1}{2}}(2\pi )^{(1N)/2}\mathrm{\Gamma }(z)\mathrm{\Gamma }(z+1/N)\mathrm{}\mathrm{\Gamma }(z+(N1)/N).$$
This implies that
(5.4)
$$\mathrm{\Gamma }_{Nn}^{}(Ns)=N^{Ns1}\mathrm{\Gamma }_n^{}(s)\mathrm{\Gamma }_n^{}(s+2/N)\mathrm{}\mathrm{\Gamma }_n^{}(s+2(N1)/N),$$
and if $`N`$ is odd then
(5.5)
$$\mathrm{\Gamma }_{Nn}^{}(Ns)=i^{(N2)(N1)/2}N^{Ns1/2}\mathrm{\Gamma }_n^{}(s)\mathrm{\Gamma }_{n+1}^{}(s+1/N)\mathrm{}\mathrm{\Gamma }_{n+N1}^{}(s+(N1)/N),$$
This can be uniformly written as
(5.6)
$$\mathrm{\Gamma }^F(\chi ^N)=C_{F,N}\chi (N^N)\mathrm{\Gamma }^F(\chi )\mathrm{\Gamma }^F(\chi \text{Nm}^{1/N})\mathrm{}\mathrm{\Gamma }^F(\chi \text{Nm}^{(N1)/N}),$$
($`N`$ is odd for $`F=`$), where Nm is the algebraic norm of $`F`$ as an extension of $``$ ($`\mathrm{Nm}(x)=x\overline{x}`$ for $``$ and $`\mathrm{Nm}(x)=x`$ for $``$), and $`C_{F,N}=N^1`$ for $``$ and $`i^{(N2)(N1)/2}N^{1/2}`$ for $``$.
### 5.3. Classification of identities with monomials.
Define the group $`X_F=X(F^{})/<\text{Nm}>`$. Let $`Div(X_F)`$ be the group of divisors on $`X_F`$. Now for $`\chi X_F`$ and a positive integer $`N`$ set $`D_{\chi ,N}=(\chi )+(\chi \text{Nm}^{1/N})+\mathrm{}+(\chi \text{Nm}^{(N1)/N})Div(X_F)`$ (here $`N`$ should be odd if $`F=`$). Set $`D_{\chi ,N}=D_{\chi ^1\nu _F,N}`$.
###### Theorem 5.2.
(i) Make the assumptions of Theorem 4.6, and suppose that the exponents $`n_i`$ are relatively prime: $`d=1`$. Then identity (4.12) holds in the weak sense for some $`C`$, $`a,b,\lambda _i,\eta _i`$ in and only in one of the following two situations:
1. $`m_i=2`$, $`m_i=n_i`$, and there exists $`\xi ,\mu _iX_F`$ such that:
(5.7)
$$D_{1,1}+\underset{i}{}D_{\mu _i,n_i}=D_{\xi ,1}.$$
2. $`m_i=0`$, $`m_i=n_i`$, and there exists $`\xi ,\mu _iX_F`$ such that:
(5.8)
$$D_{1,1}+\underset{i}{}D_{\mu _i,n_i}=D_{\xi ,1}.$$
(ii) More precisely, in any of the above two cases, one can find $`\lambda _i,\gamma X(F^{})`$, $`a,bF^{}`$ (where $`\gamma `$ is as in Theorem 4.6), so that the integral identity of Theorem 4.6 is satisfied, and $`\lambda _i\nu _F=\mu _i`$, $`\gamma =\xi `$ (resp. $`\gamma =\xi ^1`$) modulo $`\mathrm{Nm}`$ in the first (resp. second) case.
Proof. Let us sketch the proof of (i). The proof of (ii) is obtained automatically in the process of proving (i).
To prove the necessity of the conditions of (i), let us represent the elements of the group $`X_F`$ in $`X(F^{})`$ by a fundamental domain lying in the region $`\text{Re}u>>0`$. Let us write down the identity of divisors of zeros and poles for the left and right hand sides of the two Gamma function relations of Theorem 4.6 in this fundamental domain. It is easy to check that this yields exactly the two divisorial relations above.
Let us now prove the sufficiency of (i). It is not difficult to show that if the conditions of (i) are satisfied, then one can find elements $`\lambda _i,\gamma X(F^{})`$ in the cosets $`\mu _i\nu _F^1,\xi ^{\pm 1}`$ of the free cyclic group $`<\mathrm{Nm}>`$, such that the divisors of both sides of the Gamma function relations of Theorem 4.6 in $`X(F^{})`$ (not only in $`X_F`$) coincide.
By the multiplication formula (5.6), this implies that these relations hold up to a constant and an exponential factor. These factors can be removed by an appropriate choice of the constant $`C`$, and by imposing relations (5.1),(5.2) on $`a`$ and $`b`$. โ
### 5.4. Identities with monomials and exact covering systems.
Let $`(p_1,\mathrm{},p_n)`$ be positive integers whose sum equals $`p`$. According to a classical definition in elementary number theory (, problem F14) an exact covering system of type $`(p_1,\mathrm{},p_n)`$ is a representation of the group $`/p`$ as a disjoint union of cosets of groups $`/p_i`$, $`i=1,\mathrm{},n`$. It is clear that for the existence of an exact covering system of type $`(p_1,\mathrm{},p_n)`$, it is necessary that all $`p_i`$ are divisors of $`p`$. But this is not sufficient: for example, it is clear that for a covering system, $`p/p_i`$ is never relatively prime to $`p/p_j`$. In fact, there is no known simple necessary and sufficient condition for the existence of an exact covering system, and various questions about exact covering systems are the subject of an extensive theory (see , problem F14, and references therein). It is interesting, therefore, that in a special case, integral identities with monomials that we considered correspond to exact covering systems.
Consider integral identities of type (4.12) with $`n_i=2`$, $`n_k=n>0`$, and $`n_i<0`$ for $`i=1,\mathrm{},k1`$. If $`F=`$, we assume that $`n`$ is odd (this is clearly necessary for the existence of integral identities, since the multiplication formula over $``$ is valid only for multiplication by an odd number).
We will realize $`/n`$ as the set of integers $`\{0,\mathrm{},n1\}`$ and parametrize covering systems by sequences $`(p_1,\mathrm{},p_{k+1})`$ where $`p_i`$ is the biggest element of the i-th coset of the covering.
###### Theorem 5.3.
Integral identities of type (4.12) with such data for $`F=`$ are (up to rescaling $`a`$ and $`b`$) in 1-1 correspondence with exact covering systems of type $`(n_1,\mathrm{},n_{k1},1,1)`$. Moreover, the formula corresponding to the covering system $`(p_1,\mathrm{},p_{k+1})`$ has the following parameters:
(5.9)
$$\lambda _k=\mathrm{Nm}^{n1p_k},\lambda _j=\nu _F^1\mathrm{Nm}^{\frac{(p_jp_k)n_j}{n}},$$
(5.10)
$$\eta _k=\mathrm{Nm}^{p_{k+1}},\eta _j=\nu _F^1\mathrm{Nm}^{\frac{(p_{k+1}p_j)n_j}{n}}.$$
Proof. It is clear from the divisorial relation of Theorem 5.2 that to any integral identity with the given data, there corresponds a canonical exact covering system. Indeed, consider the divisor $`D_{\mu _k^1\nu _F,n}`$. According to Theorem 5.2, it is exactly covered by the divisors $`D_{\mu _i,n_i}`$, $`i=1,..,k1`$, $`D_{1,1}`$, and $`D_{\chi ^1,1}`$. Let us regard the divisor $`D_{\mu _k^1\nu _F,n}`$ as $`/n`$ by declaring $`\mu _k^1\nu _F`$ to be the unit. Then the other divisors define an exact covering of the group.
Conversely, given an exact covering system, the characters $`\lambda _i`$, $`i<k`$, are uniquely determined by $`\lambda _k`$ from the condition of cancellation of zeros and poles. Moreover, $`\lambda _k`$ itself is uniquely determined, because the Gamma-factor corresponding to the divisor $`D_{1,1}`$ does not have a shift by a character. In fact, a direct computation shows that $`\lambda _i`$ are given by the explicit formulas in the theorem. The same computation shows that $`\gamma =\nu _F\mathrm{Nm}^{(p_{k+1}p_k)/n}`$, which allows one to compute $`\eta _i`$ from the formula $`\lambda _i\eta _i\nu _F=\gamma ^{n_i}`$. The theorem is proved. โ
###### Theorem 5.4.
Integral identities corresponding to exact covering systems hold not only in the weak sense, but also in the strong sense.
Proof. It is clear from Theorem 5.3 that $`\lambda _k`$ is a monomial, so it is always strongly regular. Therefore, by Theorem 4.9, it suffices to check that there is no $`jk1`$ and singular characters $`r_j,r_k`$ such that $`(\lambda _jr_j^1)^n=(\lambda _kr_k^1)^{n_j}`$. Suppose such $`j,r_j,r_k`$ exist. Using the explicit formulas for $`\lambda _i`$, it is easy to deduce from this that $`2+p_jn(11/n_j)`$, which is impossible. The theorem is proved. โ
### 5.5. An example
Consider the simplest example of the above theory: $`k=2`$, $`n=n_2=3`$, $`n_1=1`$, $`F=`$. Applying the theorems we have proved, we obtain the following result.
###### Theorem 5.5.
Consider the distribution
(5.11)
$$\psi (ix^3/y)\lambda _2(x)\lambda _1(y).$$
(i) Such a distribution has an elementary Fourier transform in the weak sense if and only if it is one of the following six distributions:
(5.12) $`G_1=e^{ix^3/y}|y|^{1/3};G_2=e^{ix^3/y}|y|^{2/3}\text{sign}(y);G_3=e^{ix^3/y}x|y|^{4/3}\text{sign}(y);`$
$`G_4=e^{ix^3/y}x^2|y|^{5/3};G_5=e^{ix^3/y}x|y|^{2/3}\text{sign}(y);G_6=e^{ix^3/y}x^2|y|^{4/3}\text{sign}(y);`$
(ii) The Fourier transform acts on these distributions by
$$\widehat{G}_j(x,y)=\pm iG_{s(j)}(x,y/27),$$
where the sign is $`+`$ for $`j=2,3,5,6`$, and $``$ for $`1,4`$. and $`s`$ is the involution of $`\{1,2,3,4,5,6\}`$ given by $`s=(13)(45)`$. These identities hold in the strong sense.
### 5.6. Calculation of Gamma functions of prehomogeneous vector spaces over $``$ and $``$.
Now we want to study integral identities for prehomogeneous vector spaces. For this purpose, we need an explicit expression for Gamma functions of these spaces. Luckily, if $`F=`$ or $``$ then the Gamma function (up to a constant) can be found in a number of cases using Bernsteinโs polynomial. This is well known (see e.g. ), but we will give the argument for the readerโs convenience.
For simplicity we restrict ourselves to the case $`F=`$.
Consider the (multivalued) function $`f_{}()f^{s+1}`$ for $`s`$. It is easy to see that there exists a unique monic polynomial $`๐(s)`$ of degree $`D`$ such that $`f_{}()f^{s+1}=๐(s)f^s`$ (Bernsteinโs polynomial). In the case of prehomogeneous vector spaces that we are considering, Bernsteinโs polynomial was introduced by Sato in the sixties and is called โSatoโs b-functionโ (). For irreducible regular spaces, the b-functions were computed by Kimura in .
One has (for generic $`s`$):
(5.13)
$$f_{}()\lambda _{s+2,n}(f)=๐(\frac{sn}{2})\lambda _{s+1,n1}(f),\overline{f_{}()}\lambda _{s+2,n}(f)=๐(\frac{s+n}{2})\lambda _{s+1,n+1}(f).$$
Now let us find the Gamma function. We have
$$\widehat{\overline{f_{}()}h}(x)=2^D(i)^Df_{}(x)\widehat{h}(x),\widehat{f_{}()h}(x)=2^D(i)^D\overline{f_{}(x)}\widehat{h}(x).$$
Therefore, we have
$$๐(\frac{sn}{2}\frac{M}{D})\mathrm{\Gamma }^V(\lambda _{s1,n\pm 1})=2^D(i)^D\mathrm{\Gamma }^V(\lambda _{s,n}).$$
Remark. The last equation and the functional equations for the Gamma functions imply that
$$๐(sM/D)=(1)^D๐(1s).$$
This shows that the collection of roots of the polynomial $`๐(s)`$ is symmetric with respect to the point $`\frac{1}{2}(1+\frac{M}{D})`$.
From now on we will assume that $`M/D`$ is an integer, and that the roots of $`๐(s)`$ are integers or half integers. This condition is satisfied for a number of cases in . In this case, the obtained difference equation allows one to deduce a formula for the Gamma function (up to a scalar). Namely, we have
###### Proposition 5.6.
$$\mathrm{\Gamma }^V(\lambda _{s,n})=C_V\underset{j=1}{\overset{D}{}}\mathrm{\Gamma }^F(\lambda _{s+2(s_jM/D),n}),$$
where $`๐(s)=_i(s+s_i)`$, and $`C_V`$ is a constant.
Proof. Denote the proportionality coefficient between the LHS and the RHS of (5.6) by $`C_V(s,n)`$. It is obvious that this function is periodic: $`C_V(s+1,n\pm 1)=C_V(s,n)`$ and satisfies $`C_V(n,s)C_V(n,\frac{2M}{D}s)=1`$ (by virtue of the functional equation). Thus, it suffices to show that $`C_V(s,0)`$ is constant.
Since $`\mathrm{\Gamma }^V(\lambda _{s2,0})=c๐(1s)^2\mathrm{\Gamma }^V(\lambda _{s,0})`$, and $`\mathrm{\Gamma }^V(\lambda _{s,0})`$ is holomorphic for large positive $`\text{Re}(s)`$, the poles of $`\mathrm{\Gamma }^V(\lambda _{s,0})`$ can only arise at points of the form $`2(s_j+r)`$, where $`r`$ is an integer. Since $`s_j`$ are integers or half integers, this means that all the poles are integers. Similarly, because $`1/\mathrm{\Gamma }^V`$ is holomorphic for large negative $`\text{Re}(s)`$, we find that all the zeros are integers. The same clearly holds for the product of the Gamma functions on the right hand side of (5.6). Therefore, the same holds for $`C_V(s,0)`$.
Thus, we have: the meromorphic function $`h(s):=C_V(s,0)`$ is periodic with period 2, its zeros and poles are integers, and $`C_V(s,0)C_V(s,0)=1`$ (since $`2M/D`$ is an even integer). This implies that the function must have no zeros and no poles, since any zero of this function will also be its pole, and vice versa.
Given this, an asymtotic analysis for large $`s`$ (using the stationary phase approximation) shows that $`C_V(n,s)`$ is a constant. Another way to show it (knowing that $`C_V`$ has no zeros or poles) is to use the result of , which shows that $`C_V`$ is a trigonometric function. โ
### 5.7. Identities for prehomogeneous vector spaces
Let $`W`$ be a prehomogeneous vector space of dimension $`M`$ over $`F=`$, satisfying the conditions that we imposed in the previous section. Let $`f`$ be its relative invariant, of degree $`D`$. Consider the space $`V=V_1\mathrm{}V_{D1}`$, where $`V_{D1}=W`$ and $`V_j=F`$ for $`j=1,\mathrm{},D2`$. We will denote elements of $`W`$ by $`x`$ and elements of $`V_j`$ by $`t_j`$ for $`j=1,\mathrm{},D2`$. Let $`f_j=t_j`$, $`j=1,\mathrm{},D2`$, and $`f_{D1}=f`$. Let $`n_1=\mathrm{}=n_{D2}=1`$ and $`n_{D1}=D`$. We are interested in identities of type (4.24) arising in this situation.
###### Theorem 5.7.
Identities of type (4.24) for the described data, up to rescaling $`a`$ and $`b`$, correspond to functions $`\sigma :\{1,\mathrm{},D\}`$ such that $`_j(s+\sigma (j))`$ equals the b-function $`๐(s)`$ of the space $`W`$. More precisely, for any such function $`\sigma `$ the parameters $`\lambda _j,\eta _j`$ of the corresponding identity (4.24) are given by the formula
(5.14)
$$\lambda _i=\nu _F^{\sigma (i)\sigma (D1)1},iD2,\lambda _{D1}=\nu _F^{\sigma (D1)M/D},$$
(5.15)
$$\eta _i=\nu _F^{\sigma (D)\sigma (j)1},iD2,\eta _{D1}=\nu _F^{1\sigma (D)}.$$
Proof. The proof is analogous to the proof for usual monomials. โ
### 5.8. Example.
Let $`W`$ be the 27-dimensional prehomogeneous vector space over $``$ from example 6 in Section 3.4. Let $`W^{}`$ be the dual space, and $`f_{}`$ the multiplicative Legendre transform of $`f`$. The b-function of this space equals $`(s+1)(s+5)(s+9)`$ (see ). Therefore, from the previous theorem we obtain the following result.
###### Theorem 5.8.
Consider the distribution $`\psi (i\text{Re}f(x)/y)|f(x)|^p|y|^q`$ on $`W`$ ($`p,q`$).
(i) Such a distribution has an elementary Fourier transform in the weak sense if and only if $`(q,p)`$ takes one of the following six values: $`(10,8);(18,0);(6,16);`$ $`(10,0);(6,8);(14,16)`$.
(ii) The Fourier transform in the weak sense maps any of these distributions to a distribution of the same type on $`W^{}`$ (up to a scalar). The action of the Fourier transform on the parameters $`(q,p)`$ is given by the following involution:
$$(10,8)(14,16),(18,0)(6,8),(6,16)(6,16),$$
$$(10,0)(10,0).$$
The identity corresponding to $`(10,0)(10,0)`$ holds in the strong sense.
Remark. The fact that the identity corresponding to $`(10,0)(10,0)`$ holds in the strong sense is proved similarly to the case of usual monomials.
Remark. Note that the integral identities other than $`(10,0)(10,0)`$ cannot be understood in the strong sense, because the distributions we considered have poles on the divisor $`f(x)=0`$ and therefore are not well defined without a regularization. So to replace the identities we considered by identities in the strong sense one would first need to regularize one or both sides. We leave this beyond the scope of this paper.
## 6. Identities over local non-archimedean fields
### 6.1. Local constants and Gamma functions
Let $`F`$ be a local non-archimedean field. Let $`๐ชF`$ be the valuation ring, $`\pi F^{}`$ be a uniformizing element. A character $`\chi X(F^{})`$ is called unramified if $`\chi |_๐ช^{}=1`$. The local $`L`$-factors are defined as the following meromorphic functions of $`\chi X(F^{})`$:
$$L(\chi )=\{\begin{array}{cc}1\hfill & \text{ if }\chi \text{ is ramified},\hfill \\ (1\chi (\pi ))^1\hfill & \text{ if }\chi \text{ is unramified}.\hfill \end{array}$$
Note that $`L(\chi )`$ has a unique pole at the trivial character. Then we have
$$\mathrm{\Gamma }^F(\chi )=\frac{L(\chi )}{L(\chi ^1\nu _F)}ฯต(\chi ^1\nu _F)$$
where $`ฯต(\chi )=ฯต(\chi ,\psi )^{}`$ are the local constants considered by Deligne in . With our choice of the Haar measure $`dx`$ on $`F`$ we have $`ฯต(\chi )=1`$ if $`\chi `$ is unramified. If $`\chi `$ is ramified then
$$ฯต(\chi ,\psi )=_F^{}\chi ^1(x)\psi (x)๐x:=\underset{n}{}_{v(x)=n}\chi ^1(x)\psi (x)๐x.$$
Note that the functional equation (4.3) for Gamma function implies that
$$ฯต(\chi )ฯต(\chi ^1\nu _F)=\chi (1).$$
### 6.2. Identities between local constants
Recall that for a local field $`F`$ the Weil group $`W_F`$ is defined as the preimage of the subgroup generated by the Frobenius under the surjective homomorphism $`\mathrm{Gal}(\overline{F}/F)\mathrm{Gal}(\overline{k}/k)`$ where $`k`$ is the residue field of $`F`$. The local class field theory provides an isomorphism between the abelianization of the Weil group $`W_F`$ and $`F^{}`$. Following we normalize this isomorphism in such a way that uniformizing elements in $`F^{}`$ correspond to liftings of the inverse of the Frobenius. Note that for every finite separable extension $`FE`$ we have a commutative diagram
(6.1)
$$\begin{array}{ccc}W_E& \text{}& E^{}\\ \text{}& & \text{}\\ W_F& \text{}& F^{}\end{array}$$
where $`i`$ is the natural inclusion. Thus, to a character $`\chi `$ of $`F^{}`$ we can associate a one-dimensional representation of $`W_F`$. We denote by $`[\chi ]`$ the corresponding element of the representation ring $`R(W_F)`$ (of finite-dimensional continuous complex representations of $`W_F`$). If $`\lambda `$ is a character of $`F^{}`$ and $`FE`$ is a finite separable extension then $`[\lambda \mathrm{Nm}_{E/F}]=\mathrm{Res}[\lambda ]`$ where $`\mathrm{Res}:R(W_F)R(W_E)`$ is the restriction homomorphism.
The principal theorem of (Theorem 4.1) says that the map $`[\chi ]ฯต(\chi )`$ extends to a homomorphism $`Vฯต(V)`$ from $`R(W_F)`$ to $`^{}`$, such that for a finite separable extension $`E`$ of $`F`$ and for any virtual representation $`V`$ of $`W_E`$ of dimension $`0`$ one has
$$ฯต(\mathrm{Ind}_{W_E}^{W_F}V,\psi )=ฯต(V,\psi \mathrm{Tr}_{E/F})$$
(where both sides do not depend on a choice of Haar measures). If the dimension of $`V`$ is not zero then we can write
$$ฯต(\mathrm{Ind}_{W_E}^{W_F}V)=\lambda (E/F)^{dimV}ฯต(V)$$
where $`\lambda (E/F)=ฯต(\mathrm{Ind}_{W_E}^{W_F}[\nu _F^s])`$ (which does not depend on $`s`$).
On the other hand, one can extend local $`L`$-factors $`L(\chi )`$ to a homomorphism from $`R(W_F)`$ to the group of non-zero meromorphic functions on $`X(F^{})`$ by setting
$$L(V,\lambda )=det(1\mathrm{Frob}^1|_{(\lambda V)^I})^1$$
where $`IW_F`$ is the inertia subgroup (see , 3.5). Moreover, for a finite separable extension $`FE`$ and $`[V]R(W_E)`$ one has
$$L(\mathrm{Ind}_{W_E}^{W_F}[V],\lambda )=L([V],\lambda \mathrm{Nm}_{E/F})$$
(see , Prop. 3.8).
###### Proposition 6.1.
Let $`F_i`$ be finite separable extensions of a local field $`F`$, and for every $`i`$ let $`\chi _i`$ be a character of $`F_i^{}`$. Assume that we have a linear relation
$$\underset{i}{}n_i\mathrm{Ind}[\chi _i]=0$$
between the induced representations $`\mathrm{Ind}[\chi _i]`$ of $`W_F`$, where $`n_i`$. Then one has the following identity:
$$\underset{i}{}(\lambda (F_i/F)\mathrm{\Gamma }^{F_i}(\chi _i(\lambda \mathrm{Nm}_{F_i/F}))^{n_i})=1$$
for $`\lambda X(F^{})`$.
Proof. It suffices to notice that we also have the linear relation
$$\underset{i}{}n_i\mathrm{Ind}[\chi _i^1\nu _{F_i}]=0$$
(since $`\nu _{F_i}=\nu _F\mathrm{Nm}_{F_i/F}`$) and apply the above properties of $`L`$-factors and $`ฯต`$-constants. โ
Here is an example of an identity provided by the above proposition. Let $`FE`$ be a finite cyclic extension. Then for any character $`\lambda `$ of $`F^{}`$ we have the relation
$$\mathrm{Ind}_{W_E}^{W_F}[\lambda \mathrm{Nm}_{E/F}]=\underset{\chi \mathrm{Nm}_{E/F}=1}{}[\chi \lambda ]$$
where the sum is taken over all characters $`\chi `$ of $`F^{}`$ which are trivial on $`\mathrm{Nm}_{E/F}(E^{})`$ (note that this subgroup has index $`[E:F]`$ in $`F^{}`$). Hence we derive the identity
$$\lambda (E/F)\mathrm{\Gamma }^E(\lambda \mathrm{Nm}_{E/F})=\underset{\chi \mathrm{Nm}_{E/F}=1}{}\mathrm{\Gamma }^F(\chi \lambda ).$$
Remark. In the simplest case $`[E:F]=2`$ this formula appears in .
More generally, let $`FE`$ be a finite cyclic extension. For every $`n>0`$ such that $`n|[E:F]`$ let $`F_nE`$ be the cyclic subextension of degree $`n`$ over $`F`$. Let us denote by $`X_{F,E}`$ the subgroup in $`X(F^{})`$ consisting of characters of the form $`\nu _F^m\chi `$ where $`m`$, $`\chi X(F^{})`$ is trivial on $`\mathrm{Nm}_{E/F}(E^{})F^{}`$. Let $`Div(X_{F,E})`$ be the group of divisors on $`X_{F,E}`$. For every character $`\chi `$ in $`X_{F_n,E}`$ let us define the divisor $`D(\chi )Div(X_{F,E})`$ as follows:
$$D(\chi )=\underset{\lambda :\lambda \mathrm{Nm}_{F_n/F}=\chi }{}(\lambda ).$$
Note that the homomorphism
$$X_{F,E}X_{F_n,E}:\lambda \lambda \mathrm{Nm}_{F_n/F}$$
is surjective, hence the divisor $`D(\chi )`$ has degree $`n`$. As an additive character on $`F_n`$ let us take $`\psi _n=\psi \mathrm{Tr}_{F_n/F}`$. Then extending the Gamma function to divisors multiplicatively we can write
$$\lambda (F_n/F)\mathrm{\Gamma }^{F_n}(\chi (\lambda \mathrm{Nm}_{F_n/F}))=\mathrm{\Gamma }^F(\lambda D(\chi )).$$
for $`\chi X_{F_n,E}`$, $`\lambda X(F^{})`$, where for every divisor $`D`$ we denote by $`\lambda D`$ the divisor obtained from $`D`$ by applying the shift $`\mu \lambda \mu `$.
Now let $`(d_1,\mathrm{},d_k)`$ be a sequence of positive numbers dividing the degree $`[E:F]`$, and let $`(m_1,\mathrm{},m_k)`$ be a sequence of integers. For every $`i=1,\mathrm{},k`$ let $`\chi _i`$ be a character in $`X_{F_{d_i},E}`$. Assume that
$$\underset{i}{}m_iD(\chi _i)=0.$$
Then for every $`\lambda X(F^{})`$ we have
(6.2)
$$\underset{i}{}\lambda (F_{d_i}/F)^{m_i}\mathrm{\Gamma }^F(\chi _i(\lambda \mathrm{Nm}_{F_{d_i}/F}))^{m_i}=1.$$
Using the functional equation for Gamma functions we can rewrite this result slightly differently. For $`e=\pm 1`$ and a character $`\chi X_{F_n,E}`$ let us set
$$D(\chi ,e)=\{\begin{array}{cc}D(\chi ),\hfill & \text{ if }e=1,\hfill \\ D(\chi ^1\nu _{F_n}),\hfill & \text{ if }e=1.\hfill \end{array}$$
Then for a sequence $`(e_1,\mathrm{},e_k)`$, where $`e_i=\pm 1`$, the relation
$$\underset{i}{}D(\chi _i,e_i)=0$$
implies the identity
$$\underset{i}{}\mathrm{\Gamma }^F(\chi _i(\lambda ^{e_i}\mathrm{Nm}_{F_{d_i}/F}))=C\lambda (\underset{i}{}e_i)$$
for $`\lambda X(F^{})`$, where the constant $`C^{}`$ doesnโt depend on $`\lambda `$.
### 6.3. Identities for cyclic extensions
Let $`FE`$ be a cyclic extension of local fields, $`F_iE`$ be a subextension of degree $`d_i`$ over $`F`$ ($`i=1,\mathrm{},k`$). For every $`i=1,\mathrm{},k`$ let $`\chi _i`$ be a character in $`X_{F_i,E}`$, and let $`e_i`$ be either $`1`$ or $`1`$. Below we denote by $`1_F`$ the trivial character of $`F^{}`$. The following theorem follows easily from the above identities for Gamma functions and from Theorem 4.7.
###### Theorem 6.2.
In the above situation assume that one of the following identities in $`Div(X_{F,E})`$ holds:
1. $`D(1_F,1)+_iD(\chi _i,e_i)=D(\xi ,1)`$,
2. $`D(1_F,1)+_iD(\chi _i,e_i)=D(\xi ,1)`$,
for some $`\xi X_{F,E}`$. Then we have the following identity between distributions on $`_iF_i`$ in the weak sense:
$$(\underset{i}{}(\chi _i\nu _{F_i}^1)(x_i)\psi (a\underset{i}{}\mathrm{Nm}_{F_i/F}(x_i)^{e_i}))=C\underset{i}{}\eta _i(x_i)\psi (b\underset{i}{}\mathrm{Nm}_{F_i/F}(x_i)^{ee_i})$$
for some constants $`C^{}`$, $`a,bF^{}`$, where $`e=1`$ in the case 1, $`e=1`$ in the case 2,
$$\eta _i=\chi _i^1(\xi ^{ee_i}\mathrm{Nm}_{F_i/F}).$$
For example, the identity (1.2) corresponds to the following equality of divisors:
$$D(1_F,1)+D(,1)+D(\nu _E,1)=D(\nu _F,1).$$
Thus, we see that identity (1.2) holds in the weak sense.
Let us prove now that identity (1.2) in fact holds in the strong sense. We will give a sketch of the argument. First of all, the function $`\varphi _{}`$ makes sense as a distribution (i.e., does not need regularization). Since 1.2 holds in the weak sense, we have $`\widehat{\varphi _{}}=ฯต\varphi _{}+\eta `$, where $`\eta `$ is a distribution which is a sum of a distribution concentrated on the coordinate cross and a distribution whose Fourier transform is concentrated on the cross, and $`ฯต`$ is a sign. The distribution $`\eta `$ has the same homogeneity properties as $`\varphi _{}`$, and satisfies $`\widehat{\eta }=ฯต\eta `$. From this it is easy to deduce that $`\eta =c(\delta (t)ฯต\delta (x))`$. Thus, it remains to check that $`c=0`$, which can be checked by a direct calculation; for instance, one can check directly that $`\widehat{\varphi }_{}`$ is locally constant at $`x=0,t0`$. Thus, (1.2) holds in the strong sense.
### 6.4. One more identity
There are more complicated examples of identities between local constants which involve non-abelian extensions (see , section 1). Here is an example. Let $`l`$ be a positive integer, $`E`$ a Galois extension of $`F`$ with Galois group $`G`$ which is a central extension of $`(/l)^2`$ by an abelian group $`Z`$. Let $`H_1G`$ (resp. $`H_2`$) be the preimages of the first (resp. the second) factor $`/l`$ in $`(/l)^2`$. Let $`\chi _i`$ be a character of $`H_i`$ for $`i=1,2`$, such that $`\chi _1|_Z=\chi _2|_Z`$ and this character is non-trivial on $`[G,G]Z`$. Let $`F_iE`$ be the subextension of $`F`$ corresponding to the subgroup $`H_i`$. Then for $`i=1,2`$ we can consider $`\chi _i`$ as a character in $`X_{F_i,E}`$ and for every character $`\lambda X(F^{})`$ one has
$$\mathrm{Ind}_{W_{F_1}}^{W_F}\chi _1(\lambda \mathrm{Nm}_{F_1/F})=\mathrm{Ind}_{W_{F_2}}^{W_F}\chi _2(\lambda \mathrm{Nm}_{F_2/F})$$
Hence,
(6.3)
$$\lambda (F_1/F)\mathrm{\Gamma }^{F_1}(\chi _1(\lambda \mathrm{Nm}_{F_1/F}))=\lambda (F_2/F)\mathrm{\Gamma }^{F_2}(\chi _2(\lambda \mathrm{Nm}_{F_2/F}))$$
for $`\lambda X(F^{})`$. Now the following theorem follows easily from Theorem 4.7.
###### Theorem 6.3.
In the above situation we have the following identity of distributions on $`F_1\times F_2`$ in the weak sense:
$$((\chi _1\nu _{F_1}^1)(x_1)\chi _2^1(x_2)\psi (\frac{\mathrm{Nm}_{F_1/F}(x_1)}{\mathrm{Nm}_{F_2/F}(x_2)}))=C(\chi _2\nu _{F_2}^1)(x_2)\chi _1^1(1_2)\psi ((1)^l\frac{\mathrm{Nm}_{F_2/F}(x_2)}{\mathrm{Nm}_{F_1/F}(x_1)}).$$ |
warning/0003/hep-ph0003096.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Exclusive charmonium decays have been investigated within perturbative QCD by many authors, e.g. . It has been argued that the dominant dynamical mechanism is $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation into the minimum number of gluons allowed by color conservation and charge conjugation, and subsequent creation of light quark-antiquark pairs forming the final state hadrons. This factorization of long- and short-distance physics has been shown to hold in the formal limit $`m_\mathrm{c}\mathrm{}`$. The dominance of annihilation through gluons is most strikingly reflected in the narrow widths of charmonium decays into hadronic channels in a mass region where strong decays typically have widths of hundreds of MeV . Since the $`\mathrm{c}`$ and the $`\overline{\mathrm{c}}`$ quarks only annihilate if their mutual distance is less than about $`1/m_\mathrm{c}`$ (where $`m_\mathrm{c}`$ is the $`\mathrm{c}`$-quark mass) which is smaller than the non-perturbative charmonium radius, and since the average virtuality of the gluons is of the order of $`12\mathrm{GeV}^2`$ one may indeed expect perturbative QCD to be at work although corrections are presumably substantial. The charm-quark mass is too small in order to suppress power corrections decisively although it is large enough to allow perturbative QCD (pQCD) calculations. The bottomonium system, for which no exclusive hadronic decay has been observed as yet, should exhibit the pattern of perturbative predictions much cleaner.
In hard exclusive reactions higher Fock state contributions are usually suppressed by inverse powers of the hard scale, $`Q`$, appearing in the process ($`Q=m_\mathrm{c}`$ for exclusive charmonium decays), as compared to the valence Fock state contribution. Hence, higher Fock state contributions are expected to be negligible in most cases. For exclusive charmonium decays, however, the valence Fock state contributions are often suppressed for one or the other reason. In such a case higher Fock state contributions or other peculiar contributions such as power corrections or small components of the hadronic wave functions may become important. One such exception are the exclusive decays of the $`\chi _{\mathrm{cJ}}`$ mesons for which the $`\mathrm{c}\overline{\mathrm{c}}`$ pair forms a color-singlet $`\mathrm{P}`$-wave state in the valence Fock state (notation: $`\mathrm{c}\overline{\mathrm{c}}_1(^3\mathrm{P}_\mathrm{J})`$). As has been shown in Ref. the next-higher Fock state, $`\mathrm{c}\overline{\mathrm{c}}\mathrm{g}`$, where the quark-antiquark pair forms a $`\mathrm{c}\overline{\mathrm{c}}_8({}_{}{}^{3}\mathrm{S}_{1}^{})`$ state, is not suppressed as compared to the valence contribution. Therefore, the neglect of the $`\mathrm{c}\overline{\mathrm{c}}\mathrm{g}`$ contributions, which are customarily referred to as the color-octet contributions , is unjustified in the $`\chi _{\mathrm{cJ}}`$ decays. Indeed as has been discussed in Ref. for the $`PP`$ channels (where $`P`$ denotes a pseudoscalar meson) and in Ref. for their decays into baryon-antibaryon pairs the color-octet contributions are substantial and are definitely needed for obtaining agreement with experiment.
Decays into final states involving vector mesons ($`V`$) have often peculiar properties too. On general grounds, only hadronic helicity non-conserving amplitudes contribute to some of these processes. Hence, these reactions are not under control of leading-twist pQCD. A famous example of the helicity non-conserving processes is set by the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into $`\rho \pi `$. While $`J/\psi \rho \pi `$ has the largest branching ratio of all two-body hadronic $`J/\psi `$ decays, the branching ratio of $`\mathrm{\Psi }^{}\rho \pi `$ is very small . The combination of these two experimental facts forms the so-called $`\rho \pi `$ puzzle. Many of the other $`PV`$ channels behave similarly. There are many attempts to understand the $`J/\psi (\mathrm{\Psi }^{})PV`$ decays; the proposed mechanisms reach from intrinsic charm in the light mesons , to color-octet contributions , vector meson mixing , final state interactions and admixtures of a glueball nearly degenerate with the $`J/\psi `$ . Another example of a helicity non-conserving decay is set by the $`\eta _\mathrm{c}`$ decay into proton and antiproton. In Ref. it is attempted to explain this process through diquarks as quasi-elementary constituents of baryons.
In this article we are going to attempt a systematic study of the hadronic helicity non-conserving $`J/\psi (\mathrm{\Psi }^{})PV`$ decays. We will assume that, with a small probability, the charmonium possesses Fock components built from light quarks only. Through these Fock components the charmonium state decays by a soft mechanism. We will model this decay mechanism by $`J/\psi \omega \varphi `$ mixing and subsequent $`\omega `$ (or $`\varphi `$) decay into the $`PV`$ state. This, in the absence of the leading-twist perturbative QCD contribution, dominant mechanism is to be supplemented by the electromagnetic decay contribution and, depending on quantum numbers, by an anomalous, doubly Okubo-Zweig-Iizuka- (OZI-) rule violating contribution. We are not able to calculate the amplitudes of these decay processes directly but, employing flavor symmetry and mixing schemes for vector and pseudoscalar mesons, we find relations among them. With a few parameters, adjusted to experiment, we thus obtain a scheme of classification which comprises an understanding of these decay processes. The analysis of the $`J/\psi (\mathrm{\Psi }^{})PV`$ processes along these lines is partly to be considered as an update of previous work where a general parametrization of the amplitudes for $`J/\psi (\mathrm{\Psi }^{})PV`$ is fitted to experiment. We improve this parametrization by including new ideas on meson mixing and by considering recent experimental results from the BES collaboration . We also present a, partially even quantitative, physical interpretation of our parametrization which differs from those presented elsewhere . Also new in our work is the extension of the mixing approach to the $`\eta _\mathrm{c}VV`$ decays.
The paper is organized as follows: Qualitative features of charmonium decays into pairs of mesons are discussed in Sec. 2. Next, in Sec. 3, we investigate the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into the $`PV`$ channels by applying meson mixing. The mixing mechanism is also used to analyze the $`\eta _\mathrm{c}VV`$ decays (Sec. 4). In Sec. 5 we will comment on the mixing contributions to $`\mathrm{S}`$-wave charmonium decays into baryons and antibaryons briefly before we present our concluding remarks in Sec. 6. Details of our treatment of vector meson mixing are given in the Appendix.
## 2 Qualitative features of charmonium decays into meson pairs
Let us consider the $`PP`$, $`PV`$ and $`VV`$ final states. A few of these channels are strictly forbidden by angular momentum and parity conservation, see Table 1. Several other channels, characterized by non-conserved naturalness,<sup>1</sup><sup>1</sup>1 The naturalness of a meson $`i`$ having spin $`J_i`$ and parity $`P_i`$ is defined as $`\sigma _i=(1)^{J_i}P_i`$ . $`\sigma _\mathrm{c}\sigma _1\sigma _2`$, are forbidden in pQCD to leading-twist order; i.e. higher-twist or other dynamical mechanisms are at work here. This comes about for the following reasons: The helicity amplitudes for these processes are proportional to the totally antisymmetric $`ฯต`$-tensor contracted, in all possible ways, with the available Lorentz vectors, namely the two independent momenta, $`p_1`$ and $`p_2`$, the polarization vector(s) of the light vector meson(s), $`\epsilon _i`$, and, with the exception of the $`\eta _\mathrm{c}`$, the polarization vector or tensor<sup>2</sup><sup>2</sup>2 Since the polarization tensor is symmetric only one Lorentz index can be contracted with the $`ฯต`$-tensor. of the charmonium state. Now, in the rest frame of the decaying meson, the polarization vector of a helicity zero vector (or axial vector) meson can be expressed as a linear combination of the two final state momenta regardless whether or not the mass of the light vector meson is neglected. Hence, the number of independent Lorentz vectors is insufficient to contract the $`ฯต`$-tensor with the consequence of vanishing amplitudes for processes involving longitudinally polarized vector mesons. Moreover, $`\epsilon _1^{}(\pm 1)=\epsilon _2^{}(1)`$ holds in the rest frame of the decaying meson. As a consequence of these properties the only non-zero helicity amplitudes for the processes with non-conserved naturalness are those where $`\lambda _1+\lambda _20`$. In other words vector mesons are transversely polarized in reactions with non-conserved naturalness.<sup>3</sup><sup>3</sup>3 Somewhat exceptional is the $`\chi _{\mathrm{c1}}VV`$ case where one out of the three spin-1 mesons may be longitudinally polarized. We thus conclude that in these processes hadronic helicity conservation
$$\lambda _1+\lambda _2=0$$
(1)
is violated. Helicity conservation is not a consequence of a particular symmetry but is a dynamical consequence of leading-twist perturbative QCD (i.e. using leading-twist wave functions and valence Fock states only): The virtual gluons from the annihilation of the $`\mathrm{c}\overline{\mathrm{c}}`$ pair create the light, (almost) massless quarks and antiquarks in opposite helicity states. To the extent that the hadronic wave functions do not embody any non-zero orbital angular momentum components, the quark helicities sum up to their parent hadronโs helicity. Hence, the total helicity of the final state hadrons is zero. Consequently, processes that violate helicity conservation, are not governed by leading-twist pQCD. The remaining two-meson channels, marked by ticks in Table 1, are accessible to a perturbative treatment.
Next let us consider $`G`$-parity and isospin. $`G`$-parity or isospin-violating decays are not strictly forbidden since they can proceed through electromagnetic $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation and may receive contributions from the isospin-violating part of QCD. The latter contributions, being of the order of quark mass differences, seem to be small . $`G`$-parity or isospin-violating decays of $`C`$-even charmonia (e.g. $`\eta _\mathrm{c},\chi _{\mathrm{cJ}}PV`$ ($`J=1,2`$) for non-strange final state mesons) have not been observed experimentally as yet . Proceeding on the assumption that these decays are dominantly mediated by $`\mathrm{c}\overline{\mathrm{c}}2\gamma ^{}\mathrm{PV}`$, this is understandable. They are then suppressed by a factor $`(\alpha _{\mathrm{em}}/\alpha _s)^4`$ as compared to the $`G`$-parity and isospin allowed decays of the $`C`$-even charmonia and their decay widths are therefore extremely small. Channels involving strange mesons (e.g. $`KK^{}`$), should also be strongly suppressed by virtue of $`U`$-spin invariance. For $`J/\psi `$ decays the situation is different. Many $`G`$-parity violating (e.g. $`\pi ^+\pi ^{}`$) or isospin-violating (e.g. $`\omega \pi `$) decays have been observed, the experimental branching ratios being of the order of $`10^410^3`$ . As compared to $`G`$-parity and isospin allowed $`J/\psi `$ decays they are typically suppressed by factors of about $`10^210^1`$ in accordance with what is expected for an electromagnetic decay mechanism<sup>4</sup><sup>4</sup>4 A possible correction due to $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation mediated by $`\gamma ^{}g^{}g^{}`$ is ignored by us here. (see Fig. 1).
That at least two distinct dynamical mechanisms are at work in exclusive charmonium decays can be realized from a comparison of $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays. Suppose these decays are under control of a mechanism that respects QCD factorization and where the charm and anticharm quarks annihilate into gluons and/or photons at small mutual distance. The charmonium wave function is then probed at small spatial separations and, therefore, is well represented by the corresponding decay constant, $`f_{J/\psi }`$ or $`f_\mathrm{\Psi }^{}`$, which is measured in electronic decays of that charmonium state. If this short distance mechanism is responsible for the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decay into a particular two hadron channel, the ratio
$$\kappa _{12}=\frac{(\mathrm{\Psi }^{}h_1h_2)}{(J/\psi h_1h_2)}\frac{(J/\psi e^+e^{})}{(\mathrm{\Psi }^{}e^+e^{})}\left(\frac{\varrho _{12}(J/\psi )}{\varrho _{12}(\mathrm{\Psi }^{})}\right)^n$$
(2)
should be close to unity<sup>5</sup><sup>5</sup>5 In a consistent perturbative calculation to lowest order in the $`\mathrm{c}`$-quark velocity expansion one has to evaluate the hard scattering amplitude from $`2m_\mathrm{c}`$ instead from the bound state mass; the mass difference is an effect of order $`v^2`$, i.e. of the same order as the contributions from the next higher Fock states. An additional factor of $`(M_\mathrm{\Psi }^{}/M_{J/\psi })^{2n}`$ in Eq. (2) ($`n=4`$ for baryon-antibaryon channels for instance), to be found in the literature occasionally, is inconsistent in this respect. ($`n=3`$ for $`PV`$ channels and $`n=1`$ otherwise). Since $`(\mathrm{\Psi }^{}e^+e^{})/(J/\psi e^+e^{})`$ amounts to about $`14\%`$, the relation (2) is occasionally termed the $`14\%`$ rule. The phase space factor in Eq. (2) is defined by
$$\varrho _{12}(j)=\sqrt{12(m_1^2+m_2^2)/M_j^2+(m_1^2m_2^2)^2/M_j^4}$$
(3)
where $`M_j`$ is the mass of the charmonium state. It is of numerical relevance only for particles with masses above $`1\mathrm{GeV}`$. Experimental results for $`\kappa `$ are listed in Table 2; the data are taken from Refs. . For comparison we also include experimental results for the baryon-antibaryon, vectorโtensor and axial-vectorโpseudoscalar channels in the table. We observe that, within often large errors or only within bounds, $`\kappa `$ is indeed compatible with unity for the baryon-antibaryon channels, for $`b_1\pi `$, $`K^0K^0`$ and for some of the $`PV`$ channels, notably for the $`G`$-parity and isospin-violating channels $`\pi \omega `$, $`\eta \rho `$, $`\eta ^{}\rho `$. For most of the other $`PV`$ channels and, perhaps, to a lesser extent also for the vector-tensor channels, $`\kappa `$ is well below unity. We stress that the decays into a vector and a tensor meson are not forbidden by hadronic helicity conservation, but leading-twist pQCD feeds the helicity amplitude $`_{00\lambda _c}`$ only. There are many other, in general non-zero amplitudes that do not respect hadronic helicity conservation. Note that the branching ratios of $`J/\psi `$ decays into $`VT`$ are large, comparable with those for the $`PV`$ channels while those for the corresponding $`\mathrm{\Psi }^{}`$ decays are small. For the $`PP`$ channels the situation is unclear; better data are demanded. We do not include the data for multi-particle channels in the table but we note that, with the exception of the final states consisting of three pseudoscalars, $`\kappa `$ is compatible with unity for all these channels. This fact is presumably a consequence of an underlying perturbatively generated three gluon jet structure which fragments into multi-particle final states. This interpretation is supported by estimates of the total width for $`J/\psi `$ decays into light hadrons through the decay into three real gluons . In the case of the $`\mathrm{{\rm Y}}`$ decays the three jet structure is experimentally established .
The small value of $`\kappa `$ for those $`PV`$ channels that are dominated by strong interactions, arises from large $`J/\psi `$ branching ratios as an inspection of Table 2 reveals. The $`\mathrm{\Psi }^{}`$ decays into $`PV`$, on the other hand, seem to behave, at least in tendency, according to expectations based on pQCD; their branching ratios are smaller than those for the leading-twist allowed channels. Obviously, another dynamical mechanism is called for which is active for the $`J/\psi `$ decays but suppressed for the $`\mathrm{\Psi }^{}`$ ones. For this mechanism we assume that the charmonium state possesses Fock components built from light quarks only. It can then decay through these Fock components by a soft mechanism that is characteristic of OZI-rule allowed strong decays (see Fig. 2). This mechanism is obviously more peripheral than the short distance $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation; i.e. it probes the charmonium wave function at all quark-antiquark separations and feels therefore the difference between a $`1\mathrm{S}`$ and a $`2\mathrm{S}`$ radial wave function. The node in the latter is supposed to lead to a strong suppression of this mechanism in the $`\mathrm{\Psi }^{}`$ decays.<sup>6</sup><sup>6</sup>6This assumption bears a resemblance to the node effect discussed in Ref. . We model this mechanism by mixing of vector mesons, $`J/\psi \omega \varphi `$. For the mixing angle, $`\stackrel{~}{\theta }_\mathrm{c}`$, being of order $`1/m_\mathrm{c}^2`$, we take the value $`0.1^{}`$ which corresponds to a light-quark admixture to the charmonium state of order of $`10^3`$, see the Appendix. The light vector mesons, $`\omega `$ and $`\varphi `$, couple to the final state mesons through vertex functions, e.g. $`g_{\omega VP}`$. In terms of vector meson mixing the suppression of the $`\mathrm{\Psi }^{}PV`$ decays follows from substantially weaker couplings (arguments in favor of that assumption are given in Refs. ) of the radially excited $`\omega ^{}`$ and $`\varphi ^{}`$ mesons (assumed to mix with the $`\mathrm{\Psi }^{}`$) to the ground state pseudoscalar and vector mesons and, perhaps, from a smaller mixing angle. The mixing mechanism may also be at work in the vector-tensor channels (with vertex functions $`g_{VVT}`$) and may generate the small value of $`\kappa `$ there, see Table 2.
Conversely one may also consider the $`\mathrm{c}\overline{\mathrm{c}}`$ Fock components of $`\omega `$ and $`\varphi `$ mesons in the final state; their probablilities are the same as the ones for light quark components in the $`J/\psi `$ meson, see, Eq. (53) in the appendix. The initial $`\mathrm{c}\overline{\mathrm{c}}`$ pair from the $`J/\psi `$ may then feed that Fock component of the $`\omega `$ or $`\varphi `$. However, the second final state meson is to be generated by an OZI-rule violating mechanism now, and is to be considered as a higher-order correction to the leading mixing mechanism we propose. For channels like $`\omega \eta ^{}`$ one may also think of an excitation of the $`\mathrm{c}\overline{\mathrm{c}}`$ components of both the light mesons. Obviously, the probability of finding a $`\mathrm{c}\overline{\mathrm{c}}`$ pair in a light meson enters quadratically now, and this decay mechanism is suppressed by an additional power of $`1/m_c^2`$. Hence, we ignore both the above possibilities. The initial $`\mathrm{c}\overline{\mathrm{c}}`$ pair may also excite a $`\mathrm{c}\overline{\mathrm{c}}\mathrm{q}\overline{\mathrm{q}}`$ Fock component in the light (vector) meson, as suggested by Brodsky and Karliner . In that picture the $`\mathrm{q}\overline{\mathrm{q}}`$ pair from this Fock component forms the other final state meson. Of course, one expects the probability of this higher Fock state to be very small. In any case the approach of Brodsky and Karliner should be viewed as complementary to ours, since the mechanism that they propose cannot be represented by meson-mixing supplied with OZI-rule allowed decay vertices.
We note that in Refs. the strong interaction mechanism for the hadronic helicity non-conserving decays although parametrized in a similar way as we do, is interpreted differently. Thus, Chen and Braaten argue that this contribution is generated by an additional Fock state gluon (the $`\mathrm{c}\overline{\mathrm{c}}`$ pair is therefore in a color-octet state) which carries a unit of helicity from the $`J/\psi `$ to one of the light mesons without participating in the hard process. Chen and Braaten provide arguments that the suppression of the $`\mathrm{\Psi }^{}`$ decays into $`PV`$ mesons as well as into the vector-tensor channels is due to the energy gap between the mass of the $`\mathrm{\Psi }^{}`$ and the $`D\overline{D}`$ threshold. In Refs. , on the other hand, mixing of the $`J/\psi `$ with a fairly narrow $`J^{PC}=1^{}`$ glueball, nearly degenerate with the $`J/\psi `$, is assumed. The $`J/\psi `$ can then decay through the glueball while this mechanism is not available for the $`\mathrm{\Psi }^{}`$. Searches for such a glueball, performed by the BES collaboration , however turned out negative although not fully conclusive to claim the demise of the glueball interpretation. For the final state interaction mechanism advocated in Refs. the suppression of the $`\mathrm{\Psi }^{}PV`$ decays is due to a fortuitous cancellation of various contributions and, probably, does not hold for the vector-tensor channels.
The electromagnetic decay mechanism, proceeding through the subprocess $`\mathrm{c}\overline{\mathrm{c}}\gamma ^{}\mathrm{h}_1\mathrm{h}_2`$ (see Fig. 1), also probes the charmonium wave function at small spatial separation and, hence, would lead to $`\kappa 1`$ if it dominates. This is what we observe, or at least what is indicated, by the isospin-violating $`\pi \omega `$, $`\eta \rho `$ and $`\eta ^{}\rho `$ channels. For the other $`J/\psi (\mathrm{\Psi }^{})PV`$ decays a superposition of the mixing and the electromagnetic mechanism occurs. For the flavor-singlet channels one has to consider an anomalous (formally doubly OZI-rule violating) contribution too, see Fig. 3. The mixing mechanism also applies to the $`\eta _\mathrm{c}VV`$ decays with vector meson mixing replaced by that for pseudoscalar mesons .
The $`\chi _{\mathrm{cJ}}`$ may also decay through light-quark Fock components. However, in terms of meson mixing, the model would require the coupling of appropriate $`0^{++}`$, $`1^{++}`$, $`2^{++}`$ mesons to pairs of ground state mesons. These couplings are expected to be small and, therefore, the $`\chi _{\mathrm{cJ}}PP,VV`$ ($`J=0,2`$) decays seem to be dominated by perturbative contributions while the hadronic helicity non-conserving decays $`\chi _{\mathrm{c1}}VV`$ should be strongly suppressed. Indeed, the latter decays have not been observed experimentally as yet . As we already mentioned in the introduction the perturbative analysis of the $`\chi _{\mathrm{cJ}}`$ decays into the $`PP`$ channels provides results in fair agreement with experiment . In this analysis, it is however important to include the color-octet contributions, i.e. the $`\mathrm{c}\overline{\mathrm{c}}\mathrm{g}`$ Fock state of the $`\chi _{\mathrm{cJ}}`$. The contribution of the valence Fock state alone, evaluated from the asymptotic form of the pseudoscalar mesonโs light-cone wave function (as determined from the analysis of the $`P\gamma `$ transition form factors ), is substantially smaller than experiment. Aspects of the $`\eta \eta ^{}`$ mixing in these decays have been discussed in Ref. . The extension of the perturbative analysis to the $`VV`$ channels, for which first data exist , is straightforward but the required wave functions of the vector mesons are not known to a sufficient degree of accuracy at present.
## 3 Mixing mechanism and the decays $`J/\psi (\mathrm{\Psi }^{})PV`$
Let us now turn to the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into a pseudoscalar and a vector meson. According to the discussion presented in Sec. 2 we take into account the three mechanisms shown in Figs. 1, 2 and 3. Each of these contributions will be parametrized by a reduced amplitude, common to all $`PV`$ channels, multiplied by flavor factors, meson-mixing factors and flavor-symmetry corrections. Because of the lack of a compelling theory we have to treat the reduced amplitudes as free parameters to be adjusted to experiment. Despite this fact our ansatz which is structurally similar, although different in detail, to those discussed in Refs. provides a systematic description of the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into $`PV`$ mesons as we will see shortly.
To begin with we discuss the electromagnetic decay mechanism shown in Fig. 1. Decomposing the helicity amplitudes for a decay of an $`n{}_{}{}^{3}S_{1}^{}`$ charmonium state into $`PV`$ covariantly as
$`_{0\lambda _V,\lambda }(n{}_{}{}^{3}\mathrm{S}_{1}^{}VP)`$ $`=`$ $`{\displaystyle \frac{A_{PV}(n{}_{}{}^{3}\mathrm{S}_{1}^{})}{M^2(n{}_{}{}^{3}\mathrm{S}_{1}^{})}}ฯต(p_1,p_2,\epsilon _V^{}(\lambda _V),\epsilon (\lambda )),`$ (4)
we find for the electromagnetic contribution to the invariant amplitude
$`A_{PV}^{\mathrm{em}}(n{}_{}{}^{3}\mathrm{S}_{1}^{})`$ $`=`$ $`4\pi \alpha _{\mathrm{em}}e_\mathrm{c}e^{i\theta _e}f(n{}_{}{}^{3}\mathrm{S}_{1}^{})M(n{}_{}{}^{3}\mathrm{S}_{1}^{})F_{PV}(s).`$ (5)
$`f(n^3\mathrm{S}_1)`$ is the decay constant of the $`n^3\mathrm{S}_1`$ charmonium state. Its numerical value is taken from Ref. ($`f(J/\psi )=405\mathrm{MeV}`$ and $`f(\mathrm{\Psi }^{})=282\mathrm{MeV}`$). $`F_{PV}`$ is the time-like $`PV`$ transition form factor for transversely polarized vector mesons at $`s=M^2(n{}_{}{}^{3}\mathrm{S}_{1}^{})`$. We write this form factor as
$$F_{PV}(s)=c_{PV}^{\mathrm{em}}F_{\rho ^0\pi ^0}(s),$$
(6)
and treat $`F_{\rho ^0\pi ^0}(s)`$ as a free parameter ($`c_{\rho ^0\pi ^0}^{\mathrm{em}}=1`$ by definition). $`c_{PV}^{\mathrm{em}}`$ depends on the electric charges of the $`PV`$ system in units of the elementary charge $`e_0`$. We also absorb $`\eta \eta ^{}`$ and $`\omega \varphi `$ mixing factors as well as a flavor symmetry breaking correction factor, $`y^{\mathrm{em}}`$, into $`c_{PV}^{\mathrm{em}}`$. The parameter $`y^{\mathrm{em}}`$ takes into account the difference between the occurrence of strange and non-strange quarks in the form factors; its value will be estimated below. The mixing of the pseudoscalar mesons is described in the quark-flavor scheme advocated in Ref. . Vector meson mixing is treated analogously, the details are presented in the Appendix. Since the vector meson mixing angle, $`\varphi _V`$, is so small (see Eq. (52)) we put $`\mathrm{cos}\varphi _V1`$ and $`\mathrm{sin}\varphi _V\varphi _V`$. The flavor factors $`c_{PV}^{\mathrm{em}}`$ are compiled in Table 3. In Eq. (5) we also allow for a phase, $`\theta _e`$, relative to the strong interaction contributions discussed below. The transition form factors are therefore to be understood as absolute values. The invariant function $`A_{PV}`$ in Eq. (4) is normalized in such a way that the decay width reads
$$\mathrm{\Gamma }(n{}_{}{}^{3}\mathrm{S}_{1}^{}PV)=\frac{1}{96\pi }\frac{\varrho _{PV}(n{}_{}{}^{3}\mathrm{S}_{1}^{})^3}{M(n{}_{}{}^{3}\mathrm{S}_{1}^{})}|A_{PV}|^2.$$
(7)
The four isospin-violating channels now allow a simple extraction of the $`\rho ^0\pi ^0`$ form factor. From a fit to the experimental branching ratios of the isospin-violating decay channels we find
$`F_{\rho ^0\pi ^0}(M_{J/\psi }^2)`$ $`=`$ $`2710^3\mathrm{GeV}^1,`$ (8)
$`F_{\rho ^0\pi ^0}(M_\mathrm{\Psi }^{}^2)`$ $`=`$ $`1910^3\mathrm{GeV}^1.`$ (9)
The results of the fit are listed in Table 4. Note that the values (9) of the form factors are subject to errors which we estimate to 7.5% and 21% for the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ cases, respectively. The calculation of the time-like form factor, in particular in the vicinity of resonances, is a delicate issue and, in so far, reliable theoretical information on this form factor is lacking. Since the $`\rho `$ meson is transversely polarized, the form factor is under control of higher-twist contributions and should therefore fall off as $`1/s^2`$ asymptotically. Comparison of the two values in Eq. (9) rather reveals an approximate $`1/s`$ behavior in the considered region of $`s`$. A recent QCD sum rule analysis of the space-like $`\rho \pi `$ form factor yields values that are somewhat smaller than those quoted in Eq. (9). The enhancement of time-like form factors relative to space-like ones is known from the pion and nucleon cases; the disparities typically amount to about a factor of 3 in the charmonium region (see e.g. ). In any case, measurements of the $`PV`$ transition form factors in the charmonium region would be welcome. Such data would permit a critical examination of the interpretation of the electromagnetic contribution (5) (cf. also the discussion in Ref. ). In particular it might be checked whether or not the neglect of contributions from the isospin-violating part of QCD and from $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation mediated by $`\gamma ^{}g^{}g^{}`$ is indeed justified.
The $`\varphi \pi `$ channel offers direct possibility of testing $`\omega \varphi `$ mixing. As inspection of Table 4 reveals the predicted $`J/\psi \varphi \pi `$ branching ratio obtained from our value of the mixing angle (52), is well in agreement with the experimental bound. A measurement of the branching ratio for the $`\varphi \pi `$ channel would allow a direct determination of $`\varphi _V`$ through the ratio
$$\frac{(n^3S_1\varphi \pi )}{(n^3S_1\omega \pi )}=\varphi _V^2\left(\frac{\varrho _{\omega \pi }(n^3S_1)}{\varrho _{\varphi \pi }(n^3S_1)}\right)^3,$$
(10)
a relation that was proposed by Haber and Perrier long time ago.
The invariant amplitude for the strong decay mechanism (see Fig. 2) is parametrized as
$$A_{PV}^{\mathrm{mix}}(n{}_{}{}^{3}S_{1}^{})=c_{PV}^{\mathrm{mix}}g^{\mathrm{mix}}(n{}_{}{}^{3}S_{1}^{}),$$
(11)
where $`c_{PV}^{\mathrm{mix}}`$ depends on flavor symmetry breaking and meson mixing factors. As for the electromagnetic contribution we put $`c_{\rho _0\pi _0}^{\mathrm{mix}}=1`$. By virtue of the node effect (see the discussion in Sec. 2) this mechanism does not contribute to the $`\mathrm{\Psi }^{}`$ decays, i.e. we assume $`g^{\mathrm{mix}}(J/\psi )g^{\mathrm{mix}}(\mathrm{\Psi }^{})0`$ for the reduced amplitude.
The anomalous contribution, depicted in Fig. 3, can only contribute to the flavor singlet meson channels. It is assumed to proceed via a short-distance annihilation of the initial $`\mathrm{c}\overline{\mathrm{c}}`$ pair and a subsequent creation of the $`\eta `$ and $`\eta ^{}`$ states controlled by the matrix element of the topological charge density $`0|\frac{\alpha _s}{4\pi }G\stackrel{~}{G}|\eta ^({}_{}{}^{}{}_{}{}^{)}`$ which incorporates the $`\mathrm{U}(1)_\mathrm{A}`$ anomaly. As shown in Refs. the ratio of both the gluon matrix elements is given by
$$\frac{0|\frac{\alpha _s}{4\pi }G\stackrel{~}{G}|\eta }{0|\frac{\alpha _s}{4\pi }G\stackrel{~}{G}|\eta ^{}}=\mathrm{tan}\theta _8,$$
(12)
where the angle $`\theta _8`$ is related to the $`\eta \eta ^{}`$ mixing angle $`\varphi _P`$ via $`\theta _8=\varphi _P\mathrm{arctan}[\sqrt{2}/y]`$, where $`y`$ being the ratio of basic decay constants, $`f_\mathrm{q}`$ and $`f_\mathrm{s}`$ in the quark-flavor scheme . The numerical values for the mixing parameters, determined in Ref. , read $`\varphi _P=39.3^{}`$, $`\theta _8=21.2^{}`$ and $`y=0.81`$. In view of these considerations we parametrize the anomalous contribution as
$$A_{PV}^{\mathrm{anom}}(n{}_{}{}^{3}S_{1}^{})=f(n{}_{}{}^{3}S_{1}^{})c_{PV}^{\mathrm{anom}}g^{\mathrm{anom}}(n{}_{}{}^{3}S_{1}^{}),$$
(13)
and define $`c_{\omega \eta ^{}}^{\mathrm{anom}}=\sqrt{2}`$. The flavor factors $`c_{PV}^{\mathrm{anom}}`$ for the other channels can again be found in Table 3. Because of the short distance $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation mechanism $`g^{\mathrm{anom}}`$, the reduced amplitude of the anomalous contribution, should only mildly depend on the charmonium mass; we therefore assume $`g^{\mathrm{anom}}(J/\psi )=g^{\mathrm{anom}}(\mathrm{\Psi }^{})`$ for simplicity.
A fit to the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decay widths into $`\rho \pi `$, $`\omega \eta `$ and $`\omega \eta ^{}`$ provides the results shown in Table 4 for the following values of the parameters
$`g^{\mathrm{mix}}(J/\psi )=0.020\mathrm{GeV},`$ (14)
$`g^{\mathrm{anom}}=0.014,`$ (15)
$`\theta _e=78^{}.`$ (16)
The value of the relative angle between the electromagnetic and the strong contributions is very similar to that found by Bramon et al. . Relative phases with values close to $`90^{}`$ between different decay amplitudes have also been observed in other approaches to $`J/\psi `$ ($`\psi ^{}`$) decays . Concerning the relative size of contributions from the three individual decay mechanisms (electromagnetic, mixing, anomalous), the result strongly depends on the considered decay channel. For instance, in the case of $`J/\psi \rho \pi `$ the ratio of $`|A^{\mathrm{em}}|/|A^{\mathrm{mix}}|`$ is only $`0.1`$ whereas for $`J/\psi K^{}{}_{}{}^{0}\overline{K}_{}^{0}`$ the same ratio is about $`0.3`$. The anomalous contribution tends to interfere with the mixing contribution destructively and is in some cases very important. This is most drastically seen in the channel $`J/\psi \varphi \eta ^{}`$ where the ratio $`|A^{\mathrm{em}}|/|A^{\mathrm{mix}}+A^{\mathrm{anom}}|`$ is about $`1.1`$.
The decays into $`KK^{}`$ and $`\varphi \eta (^{})`$ involve strange quarks which necessitates the consideration of flavor symmetry breaking. In the case of the electromagnetic contribution one has to pay attention to the fact that the $`PV`$ transition form factors are controlled by higher-twist contributions since the vector mesons are transversely polarized. Hence, the form factors are proportional to a mass scale which, as shown by Efremov and Teryaev , is of the order of the hadron masses and not, as one may naively expect, set by current-quark masses. Modelling this mass scale by appropriate effective constituent-quark masses ($`\widehat{m}`$), we estimate $`y^{\mathrm{em}}\widehat{m}_\mathrm{s}/\widehat{m}_\mathrm{q}1.3`$. For $`\varphi \eta `$ and $`\varphi \eta ^{}`$ channels the factor should appear quadratically. For the mixing contribution strange quarks come either from $`J/\psi \varphi `$ mixing, and are therefore $`\mathrm{cot}\stackrel{~}{\theta }_y`$, or by a soft creation of a $`\mathrm{s}\overline{\mathrm{s}}`$ pair out of the vacuum. This creation process is suppressed by a factor that is less than unity, relative to the creation of a non-strange quark-antiquark pair. In accordance with the flavor symmetry breaking in vector meson mixing, see the Appendix, we identify this factor with $`\stackrel{~}{y}`$ and use the value 0.8 for it. Note that in our mixing scheme $`\sqrt{2}\mathrm{cot}\stackrel{~}{\theta }_y=\stackrel{~}{y}`$, see Eq. (45). Therefore, the pattern of $`\mathrm{SU}(3)_\mathrm{F}`$ breaking is similar for the mixing and the electromagnetic contributions. With these values for the flavor symmetry factors and the other already determined parameters at hand we are in the position to predict the branching ratios for the decay channels involving strange quarks. The results are also shown in Table 4. The quality of the predictions for the entire set of $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into a $`PV`$ pair is good. For the $`J/\psi `$ decays the quality is similar to that obtained in Ref. although we have a smaller number of adjustable parameters.
In our approach the $`\mathrm{\Psi }^{}`$ decays into the $`PV`$ pair electromagnetically with the exception of the flavor-singlet channels where the anomalous mechanism contributes as well. In so far our analysis differs from that advocated by Tuan who extended the analysis in Ref. to the case of the $`\mathrm{\Psi }^{}`$. Tuan assumes the electromagnetic and the strong contributions to cancel approximately in the $`\mathrm{\Psi }^{}\rho \pi `$ channel. This assumption leads to substantially smaller branching ratios for the $`\mathrm{\Psi }^{}\rho \pi `$ and $`\mathrm{\Psi }^{}K^+K^{}`$ channels than we obtain. Since for both channels the BES collaboration only provides bounds the issue which of the contending models is correct, cannot be settled as yet.
Let us now dwell upon the compatibility of the fit value for $`g(J/\psi )`$ with the concept of $`J/\psi \omega \varphi `$ mixing. In this approach one would write the reduced amplitude as
$$g^{\mathrm{mix}}(J/\psi )M_{J/\psi }^2\stackrel{~}{\theta }_\mathrm{c}\mathrm{sin}\stackrel{~}{\theta }_yg_{\omega \rho ^0\pi ^0}(s=M_{J/\psi }^2).$$
(17)
The $`\omega \rho ^0\pi ^0`$ vertex function at $`s=0`$ can be estimated from chiral anomaly predictions for the coupling constants $`g_{P\gamma \gamma }`$ and vector meson dominance<sup>7</sup><sup>7</sup>7This approach has been used to predict the radiative $`PV`$ transition, see and references therein. Kramer, Palmer and Pinsky investigated the $`\omega \rho ^0\pi ^0`$ vertex on the basis of effective Lagrangians. Their numerical results are in agreement with Eq. (18). They also found $`|g_{\omega ^{}\rho ^0\pi ^0}(s=0)||g_{\omega \rho ^0\pi ^0}(s=0)|`$ which is in line with our assumption $`g^{\mathrm{mix}}(\mathrm{\Psi }^{})g^{\mathrm{mix}}(J/\psi )`$.:
$$g_{\omega \rho ^0\pi ^0}(s=0)\frac{6M_\rho M_\omega }{4\pi ^2\sqrt{2}f_\pi f_\rho f_\omega }.$$
(18)
Using our values for the $`J/\psi \omega `$ mixing parameters (52) and a monopole form for the $`s`$ dependence of the vertex function, $`\mathrm{\Lambda }^2/(M_{J/\psi }^2\mathrm{\Lambda }^2)`$, we find that the fit value of $`g^{\mathrm{mix}}(J/\psi )`$ (16) is reproduced for $`\mathrm{\Lambda }1\mathrm{GeV}`$. With regard to all the uncertainties encountered in an estimate like this, a value of about 1 GeV for $`\mathrm{\Lambda }`$ appears reasonable and is compatible with the typical inverse radius of a light hadronic system. Thus, we conclude that the small probability of the light-quark Fock components of the $`J/\psi `$ as estimated by $`J/\psi \omega \varphi `$ mixing, suffices to generate the large $`J/\psi PV`$ branching ratios seen in experiment since it is overcompensated by the very large soft $`\omega ,\varphi PV`$ transition.
## 4 Decays $`\eta _\mathrm{c}VV`$
We write the helicity amplitudes for the $`\eta _\mathrm{c}VV`$ decays in a similar fashion as in Eq. (4) and parametrize the invariant amplitude analogously to Eq. (11) as
$$A_{VV}(\eta _\mathrm{c})=c_{VV}^{\mathrm{mix}}g^{\mathrm{mix}}(\eta _\mathrm{c}),$$
(19)
with the decay width given by
$$\mathrm{\Gamma }(\eta _cVV)=\frac{1}{32\pi }\frac{\varrho _{VV}(\eta _c)}{M_{\eta _c}}|A_{VV}|^2.$$
(20)
We assume that the $`\eta _\mathrm{c}`$ decay is mediated by the mixing of the $`\eta _\mathrm{c}`$ meson with the $`\eta `$ and $`\eta ^{}`$. The $`\eta \eta ^{}`$ mixing parameters have already been quoted. The mixing mechanism leads to the flavor factors $`c_{VV}^{\mathrm{mix}}`$ shown in Table 5 ($`c_{\rho ^0\rho ^0}^{\mathrm{mix}}=1`$). Electromagnetic contributions to $`\eta _\mathrm{c}`$ decays can only proceed via two photons and are thus negligible (see also footnote 4). Also a doubly OZI-rule violating decay mechanism should be strongly suppressed here since the production of light vector mesons through gluons is not enhanced by the $`U(1)_A`$ anomaly. The possibility of a decay through the $`\mathrm{c}\overline{\mathrm{c}}`$ Fock component of the light mesons is, as in the case of the $`J/\psi `$, ignored, see the discussion in Sec. 2.
A fit to the experimental branching ratios <sup>8</sup><sup>8</sup>8Since the two results for the $`\varphi \varphi `$ channel quoted in Refs. are not compatible with each other, we have doubled the respective errors in the fit. yields $`g^{\mathrm{mix}}(\eta _\mathrm{c})=0.26`$GeV. The corresponding results are listed in Table 5. The strength of $`g^{\mathrm{mix}}(\eta _\mathrm{c})`$ is compatible with the concept of mixing. To see this we first note that here the same vertex function, $`g_{PVV},`$ enters as in the $`J/\psi PV`$ decays in the vector meson mixing approach. Analogously to Eqs. (17,18) the reduced amplitude reads
$$g^{\mathrm{mix}}(\eta _\mathrm{c})M_{\eta _\mathrm{c}}^2|\theta _\mathrm{c}|\mathrm{sin}(\varphi _P\theta _8)g_{\eta _q\rho ^0\rho ^0}(s=M_{\eta _\mathrm{c}}^2),$$
(21)
and the vertex function at $`s=0`$ is estimated by
$$g_{\eta _q\rho ^0\rho ^0}(s=0)\frac{6M_\rho ^2}{4\pi ^2\sqrt{2}f_qf_\rho ^2}.$$
(22)
Inserting the value $`1.0^{}`$ for the mixing angle $`\theta _c`$ that controls the light quark admixture to the $`\eta _c`$ as well as $`f_q=1.07f_\pi `$ and assuming again a monopole behavior of the vertex function the value $`g^{\mathrm{mix}}(\eta _\mathrm{c})=0.26\mathrm{GeV}`$ is reproduced for $`\mathrm{\Lambda }1.3`$ GeV. This indicates a milder $`s`$ dependence of the vertex function in Eq. (21) than in Eq. (17). We stress that the mixing approach explains the order of magnitude of the $`\eta _\mathrm{c}VV`$ decay widths correctly. This is, to our opinion, a highly non-trivial fact. In tendency our results agree better with the experimental results quoted by the DM-2 collaboration than with the ones found by MARK-III . Better data for the $`\eta _\mathrm{c}VV`$ channels are obviously needed. In analogy to the $`\psi ^{}PV`$ decays the $`\eta _\mathrm{c}^{}VV`$ decays are expected to be strongly suppressed.
## 5 Comments on decays into baryon-antibaryon pairs
A perturbative calculation of the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into baryon-antibaryon pairs has been carried through in Ref. using the light-cone wave function for baryons proposed in Ref. . In contrast to the $`\chi _{\mathrm{cJ}}`$ decays here it suffices to consider the color-singlet contributions only. The color-octet contributions are suppressed by powers of $`1/m_\mathrm{c}`$ and start at the order of the charm-quark velocity, $`v`$, squared. Moreover, there is no obvious enhancement of the corresponding hard scattering amplitudes which appear with at least the same power of $`\alpha _s`$ as the valence Fock state contributions. Thus, despite the fact that $`m_\mathrm{c}`$ is not very large and $`v`$ not small ($`v^20.3`$), it seems reasonable to expect small higher Fock state contributions to both these decays. One may, however, wonder what the strength of the mixing mechanism is. Since this mechanism is responsible for the large $`PV`$ branching ratios (see Sec. 2) it may also contribute to the baryon-antibaryon channels substantially and spoil the results presented in Ref. . An estimate of the mixing contribution in analogy to Eqs. (17,18), using, for instance, the $`\omega `$-nucleon coupling constants $`g_{V\omega p\overline{p}}^2/4\pi =8.1\pm 1.5`$ and $`g_{T\omega p\overline{p}}^2/4\pi =0.14\pm 0.2`$ , reveals that it amounts to less than $`20\%`$ of the experimental width if the $`s`$ dependence of the $`\omega p\overline{p}`$ vertex function is at least as strong as that of the $`\omega \rho \pi `$ one. Since the $`s`$ dependence of a baryon vertex is likely to be stronger than that of a meson vertex the mixing contribution to the baryon-antibaryon channels is probably very small. Therefore the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into baryon-antibaryon pairs seem to be under control of perturbative QCD. Indeed, the order $`v^0`$ perturbative results obtained in Ref. agree well with experiment . As we emphasized in Sec. 2 the dominance of the perturbative contribution is further supported by the fact that the $`\mathrm{\Psi }^{}J/\psi `$ ratio $`\kappa _{B\overline{B}}`$, defined in Eq. (2), is in agreement with unity (see Table 2) as is indicative for short-distance $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation.
The hadronic helicity non-conserving decay $`\eta _\mathrm{c}p\overline{p}`$ may be estimated through $`\eta _\mathrm{c}\eta \eta ^{}`$ mixing. Using the coupling constant $`g_{\eta N\overline{N}}^2/4\pi 1`$ and $`g_{\eta ^{}N\overline{N}}0`$ and equating the $`s`$ dependence of the $`\eta N\overline{N}`$ vertex with that of the $`\eta VV`$ one as determined from the fit discussed in Sec. 4, we find $`(\eta _\mathrm{c}p\overline{p})0.510^3`$. Comparing with the experimental value of $`(1.2\pm 0.4)10^3`$ we are tempted to conclude that the mixing mechanism also controls the $`\eta _\mathrm{c}p\overline{p}`$ decays.
## 6 Concluding remarks
Exclusive charmonium decays constitute a laboratory for investigating power corrections and higher Fock state contributions as well as for studying the interplay of pQCD and soft mechanisms. That there are two distinct dynamical mechanisms at work can most easily be seen by comparison of $`J/\psi PV`$ and $`\mathrm{\Psi }^{}PV`$ decays. Their scaling with the charmonium decay constants is indicative for a short-distance mechanism (as, for instance, $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation into virtual gluons or photons) while a clear deviation from that scaling (as is seen for the helicity non-conserving $`PV`$ channels) signals the prominent role of a more peripheral soft mechanism. Several of the charmonium decays into pairs of pseudoscalar and/or vector mesons are forbidden by spin and parity invariance. Others, for instance the decays of the $`C`$-even charmonia into $`PV`$ channels, are under control of the electromagnetic decay mechanism because either $`G`$-parity or isospin conservation is violated for the channels involving non-strange mesons. Therefore, only the following classes of decays are actually possible: the helicity non-conserving $`J/\psi (\mathrm{\Psi }^{})PV`$ and $`\eta _\mathrm{c}VV`$ decays which are not dominated by leading-twist pQCD; and the $`\chi _{\mathrm{cJ}}PP,VV`$ ($`J=0,2`$) decays which are accessible to a perturbative treatment and the $`J/\psi (\mathrm{\Psi }^{})PP,VV`$ decays. The latter decays are of electromagnetic nature, the amplitudes being related to the time-like $`PP`$ or $`VV`$ electromagnetic form factors at $`s=M^2(n^3\mathrm{S}_1)`$. One should, however, be aware of corrections to the electromagnetic decay mechanism from the isospin-violating part of pQCD. In principle, allowed by strong interactions are also the helicity non-conserving $`\chi _{\mathrm{c1}}VV`$ decays. We however expect very small branching ratios for these channels. The decays of the $`C`$-even charmonia into the $`PV`$ channels proceed through $`\mathrm{c}\overline{\mathrm{c}}`$ annihilation into two photons and are therefore strongly suppressed. In fact these decays as well as the $`\chi _{\mathrm{c1}}VV`$ have not been observed in experiment as yet .
The perturbative analysis of the $`\chi _{\mathrm{cJ}}PP`$ decays has been carried through in Refs. and the results have been found to agree fairly well with experiment although, with regard to the recent BES data , some readjustment of the wave function parameters seems to be advisable. The $`VV`$ channels, on the other hand, have not been analyzed as yet. Note that the required light-cone wave functions of the vector mesons are not known to a sufficient degree of accuracy.
In the present paper we focus on the $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays into the $`PV`$ channels as well as the $`\eta _\mathrm{c}VV`$ decays. We assume that in these cases the charmonium state decays dominantly by a soft mechanism through Fock components built up from light quarks only. This mechanism probes all quark-antiquark separations in the charmonium wave functions and therefore feels the difference between a $`1\mathrm{S}`$ and $`2\mathrm{S}`$-wave function. We model this mechanism by $`J/\psi \omega \varphi `$ and $`\eta _\mathrm{c}\eta \eta ^{}`$ mixing, respectively, and argue that this contribution is negligible for $`\mathrm{\Psi }^{}`$ decays. The light vector and pseudoscalar mesons couple through vertex functions to the final state mesons. With a few parameters adjusted to experiment we find fair agreement with the many channels measured for $`J/\psi (\mathrm{\Psi }^{})PV`$ and $`\eta _\mathrm{c}VV`$ decays. Most of the parameters have a simple, direct physical interpretation and their magnitudes can be estimated from physical considerations. Data of improved quality, in particular for the $`\mathrm{\Psi }^{}`$ and $`\eta _\mathrm{c}`$ decays are required for a more stringent test of our approach. A richer database for the vector-tensor channels in $`J/\psi `$ and $`\mathrm{\Psi }^{}`$ decays would also allow the extension of our mixing approach to this class of reactions.
## Acknowledgments
We thank Thomas Teubner for a critical reading of the manuscript.
## Appendix: Vector meson mixing
In order to estimate the strength of the light-quark Fock components in the $`J/\psi `$ we consider $`J/\psi \omega \varphi `$ mixing in parallel to the mixing of the pseudoscalar mesons. As in Refs. (see also ) we start from the quark-flavor basis formed by the three states $`\omega _\mathrm{q}`$, $`\omega _\mathrm{s}`$ and $`\omega _\mathrm{c}`$ which are assumed to have the parton decompositions
$`\omega _\mathrm{q}`$ $`=`$ $`\mathrm{\Psi }_\mathrm{q}\mathrm{u}\overline{\mathrm{u}}+\mathrm{d}\overline{\mathrm{d}}/\sqrt{2}+\mathrm{},`$
$`\omega _\mathrm{s}`$ $`=`$ $`\mathrm{\Psi }_\mathrm{s}\mathrm{s}\overline{\mathrm{s}}+\mathrm{},`$ (23)
$`\omega _\mathrm{c}`$ $`=`$ $`\mathrm{\Psi }_\mathrm{c}\mathrm{c}\overline{\mathrm{c}}+\mathrm{}`$
where the ellipses stand for higher Fock states. The physical meson states are related to the basis (Appendix: Vector meson mixing) by an orthogonal transformation
$`\left(\begin{array}{c}\omega \\ \varphi \\ J/\psi \end{array}\right)=U(\varphi _V,\stackrel{~}{\theta }_y,\stackrel{~}{\theta }_\mathrm{c})\left(\begin{array}{c}\omega _\mathrm{q}\\ \omega _\mathrm{s}\\ \omega _\mathrm{c}\end{array}\right)`$ (30)
where
$`U`$ $`=`$ $`\left(\begin{array}{ccc}1& \varphi _V& \stackrel{~}{\theta }_\mathrm{c}\mathrm{sin}\stackrel{~}{\theta }_y\\ \varphi _V& 1& \stackrel{~}{\theta }_\mathrm{c}\mathrm{cos}\stackrel{~}{\theta }_y\\ \stackrel{~}{\theta }_\mathrm{c}\mathrm{sin}\stackrel{~}{\theta }_y& \stackrel{~}{\theta }_\mathrm{c}\mathrm{cos}\stackrel{~}{\theta }_y& 1\end{array}\right),`$ (34)
($`U^{}U=1+๐ช(\stackrel{~}{\theta }_\mathrm{c}^2,\varphi _V^2,\stackrel{~}{\theta }_\mathrm{c}\varphi _V)`$). In Eq. (34) $`\stackrel{~}{\theta }_\mathrm{c}`$ and $`\varphi _V`$ are considered as small quantities. $`\stackrel{~}{\theta }_\mathrm{c}`$ is small since the mixing between the light and the heavy sector is an effect of $`๐ช(1/m_\mathrm{c}^2)`$ while the smallness of $`\varphi _V`$ follows from the well-known fact that the $`\omega `$ and the $`\varphi `$ are nearly ideally mixed .
The mass matrix in the quark-flavor basis is related to the (diagonal) mass matrix for the physical mesons by
$`_{\mathrm{qsc}}^2`$ $`=`$ $`U^{}(\varphi _V,\stackrel{~}{\theta }_y,\stackrel{~}{\theta }_\mathrm{c})\mathrm{diag}[M_\omega ^2,M_\varphi ^2,M_{J/\psi }^2]U(\varphi _V,\stackrel{~}{\theta }_y,\stackrel{~}{\theta }_\mathrm{c}).`$ (35)
Following our previous work on the pseudoscalar mesons , we parametrize the mass matrix as
$`_{\mathrm{qsc}}^2`$ $`=`$ $`\left(\begin{array}{ccc}m_{qq}^2+2\stackrel{~}{a}^2& \stackrel{~}{y}\sqrt{2}\stackrel{~}{a}^2& \stackrel{~}{z}\sqrt{2}\stackrel{~}{a}^2\\ \stackrel{~}{y}\sqrt{2}\stackrel{~}{a}^2& m_{ss}^2+\stackrel{~}{y}^2\stackrel{~}{a}^2& \stackrel{~}{y}\stackrel{~}{z}\stackrel{~}{a}^2\\ \stackrel{~}{z}\sqrt{2}\stackrel{~}{a}^2& \stackrel{~}{y}\stackrel{~}{z}\stackrel{~}{a}^2& m_{cc}^2\end{array}\right),`$ (39)
where $`\stackrel{~}{y}`$ and $`\stackrel{~}{z}`$ allow for flavor symmetry breaking effects in the otherwise โdemocraticโ mixing matrix whose strength, $`\stackrel{~}{a}^2`$, being controlled by quark-annihilation processes.
Comparing the expressions (35) and (39) we find the following six relations between the nine parameters
$`m_{qq}^2+2\stackrel{~}{a}^2=M_\omega ^2,`$ (40)
$`m_{ss}^2+\stackrel{~}{y}^2\stackrel{~}{a}^2=M_\varphi ^2,`$ (41)
$`m_{cc}^2=M_{J/\psi }^2,`$ (42)
$`\sqrt{2}\stackrel{~}{y}\stackrel{~}{a}^2=(M_\varphi ^2m_\omega ^2)\varphi _V,`$ (43)
$`\mathrm{tan}\stackrel{~}{\theta }_y=\sqrt{2}/\stackrel{~}{y},`$ (44)
$`\stackrel{~}{\theta }_\mathrm{c}=\sqrt{2+\stackrel{~}{y}^2}\stackrel{~}{z}\stackrel{~}{a}^2/M_{J/\psi }^2.`$ (45)
One way to utilize these relations is to make use of flavor symmetry in order to fix the quark mass terms. Thus, one may use
$$m_{qq}^2M_\rho ^2,m_{ss}^22M_K^{}^2M_\rho ^2,$$
(46)
in analogy to the pseudoscalar meson case . This would allow the determination of the mixing parameters with the exception of $`\stackrel{~}{z}`$. However, we dismiss this method since it leads to values for the parameters which are instable under small corrections to the relations (46).
An alternative method that leads to robust values for the mixing parameters, is to take the mixing angle $`\varphi _V`$ from experiment and to fix the parameters controlling flavor symmetry breaking in analogy to the pseudoscalar case by
$$\stackrel{~}{y}=f_\omega /f_\varphi ,\stackrel{~}{z}=f_\omega /f_{J/\psi }.$$
(47)
This is an assumption here in contrast to the pseudoscalar case where the divergences of the axial vector currents, embodying the $`U(1)_A`$ anomaly, relate the flavor symmetry breaking parameters to the meson decay constants. From the values of the decay constants quoted in Ref. , we obtain $`\stackrel{~}{y}=0.80\pm 0.04`$ and $`\stackrel{~}{z}=0.48\pm 0.02`$. The mixing ansatz (30) directly relates the ratio of the widths for the radiative $`\varphi \pi ^0`$ and $`\omega \pi ^0`$ transitions to the mixing angle $`\varphi _V`$
$$\frac{\mathrm{\Gamma }(\varphi \pi ^0\gamma )}{\mathrm{\Gamma }(\omega \pi ^0\gamma )}=\varphi _V^2\left[\frac{M_\varphi }{M_\omega }\frac{\varrho _{\pi \gamma }(\varphi )}{\varrho _{\pi \gamma }(\omega )}\right]^3.$$
(48)
From the experimental decay widths we find
$$\varphi _V=3.4^{}\pm 0.2^{},$$
(49)
where the sign is chosen according to Eq. (45). This result is consistent with other determinations of that angle .
Using Eqs. (47,49) and of course the masses of the physical vector mesons as input to Eq. (45), we obtain for the other mixing parameters
$`\stackrel{~}{a}^2=0.022\pm 0.002\mathrm{GeV}^2,`$ (50)
$`\stackrel{~}{\theta }_y=59.8^{}\pm 1.3^{},`$ (51)
$`\stackrel{~}{\theta }_\mathrm{c}=0.10^{}\pm 0.01^{}.`$ (52)
The mixing strength, $`\stackrel{~}{a}^2`$ is much smaller than in the case of the pseudoscalar mesons where it is dominated by the $`U(1)_A`$ anomaly. For completeness we note that an evaluation of the quark mass terms from these values for the mixing parameters leads to values which merely deviate by about $`2\%`$ from those obtained via the flavor symmetry relations (46).
Our approach leads to the following quark content of the physical mesons
$`|\omega `$ $`=`$ $`|\omega _\mathrm{q}0.060|\omega _\mathrm{s}1.510^3|\omega _\mathrm{c},`$
$`|\varphi `$ $`=`$ $`|\omega _\mathrm{s}+0.060|\omega _\mathrm{q}0.910^3|\omega _\mathrm{c},`$
$`|J/\psi `$ $`=`$ $`|\omega _\mathrm{c}+1.510^3|\omega _\mathrm{q}+0.910^3|\omega _\mathrm{s}.`$ (53) |
warning/0003/cond-mat0003084.html | ar5iv | text | # Moment free energies for polydisperse systems
## I Introduction
The thermodynamics of mixtures of several chemical species is, since Gibbs, a well established subject (see e.g. ). But many systems arising in nature and in industry contain, for practical purposes, an infinite number of distinct, though similar, chemical species. Often these can be classified by a parameter, $`\sigma `$, say, which could be the chain length in a polymeric system, or the particle size in a colloid; both are routinely treated as continuous variables. In other cases (see e.g. ) $`\sigma `$ is instead a parameter distinguishing species of continuously varying chemical properties. In the most general case, several attributes may be required to distinguish the various particle species in the system (such as length and chemical composition in length-polydisperse random copolymers) and $`\sigma `$ is then a collection of parameters . The thermodynamics of polydisperse systems (as defined above) is of crucial interest to wide areas of science and technology; it is sometimes also referred to as โcontinuous thermodynamicsโ (see e.g. ).
Standard thermodynamic procedures for constructing phase equilibria in a system of volume $`V`$ containing $`M`$ different species can be understood geometrically in terms of a free energy surface $`f(\rho _j)`$ (with $`f=F/V`$, and $`F`$ the Helmoltz free energy) in the $`M`$-dimensional space of number densities $`\rho _j`$. Tangent planes to $`f`$ define regions of coexistence, within which the free energy of the system is lowered by phase separation. The volumes of coexisting phases follow from the well-known โlever ruleโ. Here โsurfaceโ and โplaneโ are used loosely, to denote manifolds of appropriate dimension. This procedure becomes unmanageable, both conceptually and numerically, in the limit $`M\mathrm{}`$ which formally defines a polydisperse system. There is now a separate conserved density $`\rho (\sigma )`$ for each value of $`\sigma `$; $`\rho (\sigma )`$ is in fact a density distribution and the overall number density of particles is written as $`\rho =๐\sigma \rho (\sigma )`$. The free energy surface is $`f=f[\rho (\sigma )]`$ (a functional) which resides in an infinite-dimensional space. Gibbsโ rule allows the coexistence of arbitrarily many thermodynamic phases.
To make phase equilibria in polydisperse systems more accessible to both computation and physical intuition, it is clearly helpful to reduce the dimensionality of the problem. Theoretical work in this area has made significant progress in finding simplified forms of the conditions for phase coexistence, thus making these more numerically tractable . Our aim is to achieve a similar reduction in dimensionality on a higher level, that of the free energy itself. We show that, for a large class of models, it is possible to construct a reduced free energy depending on only a small number of variables. From this, meaningful information on phase equilibria can be extracted by the usual tangent plane procedure, with obvious benefits for a more intuitive understanding of the phase behavior of polydisperse systems. As an important side effect, this procedure also leads to robust algorithms for calculating polydisperse phase equilibria numerically. In particular, such algorithms can handle coexistence of more than two phases with relative ease compared to those used previously.
A clue to the choice of independent variables for such a reduced free energy comes from the representation of experimental data. We recall first the definition of a โcloud pointโ (see e.g. ): This is the point at which, for a system with a given density distribution $`\rho (\sigma )`$, phase separation first occurs as the temperature $`T`$ or another external control parameter is varied. The corresponding incipient phase is called the โshadowโ. Now consider diluting or concentrating the system, i.e., varying its overall density $`\rho `$ while maintaining a fixed โshapeโ of polydispersity $`n(\sigma )=\rho (\sigma )/\rho `$. We will find it useful later to refer to the collection of all systems $`\rho (\sigma )=\rho n(\sigma )`$ which can be obtained by this process as a โdilution lineโ (in the space of all density distributions $`\rho (\sigma )`$). Plotting the cloud point temperature $`T`$ versus density $`\rho `$ then defines the cloud point curve (CPC), while plotting $`T`$ versus the density of the shadow gives the shadow curve . The differences in the shapes $`n(\sigma )`$ of the density distributions in the different phases are hidden in this representation; only the overall densities $`\rho `$ in each phase are tracked. The density $`\rho `$ is a particular moment of $`\rho (\sigma )`$ (the zeroth one); higher order moments would be given by $`๐\sigma \sigma ^i\rho (\sigma )`$. Generalizing slightly, this suggests that our reduced free energy should depend on several (generalized) moment densities $`\rho _i=๐\sigma w_i(\sigma )\rho (\sigma )`$, defined by appropriate (linearly independent) weight functions $`w_i(\sigma )`$. Ordinary power-law moment densities are included as the special case $`w_i(\sigma )=\sigma ^i`$. Whatever the choice of moments, we insist on $`w_0=1`$ so that in all cases the zeroth moment density $`\rho _0`$ coincides with the overall number density $`\rho `$.
As an illustration, consider the simplest imaginable case where the true free energy $`f`$ already has the required form, i.e., it depends only on a finite set of $`K`$ moment densities of the density distribution:
$$f=f(\rho _i),i=1\mathrm{}K\mathrm{o}ri=0\mathrm{}K1$$
(1)
where the range of $`i`$ depends on whether $`\rho _0\rho `$ is among the $`K`$ moment densities on which $`f`$ depends. (From now on, we use the notation $`i=1\mathrm{}K`$ inclusively, to cover both possibilities.) In coexisting phases one demands equality of particle chemical potentials, defined as $`\mu (\sigma )=\delta f/\delta \rho (\sigma )`$, for all $`\sigma `$. Because
$`\mu (\sigma ){\displaystyle \frac{\delta f}{\delta \rho (\sigma )}}={\displaystyle \underset{i}{}}{\displaystyle \frac{f}{\rho _i}}w_i(\sigma )={\displaystyle \underset{i}{}}\mu _iw_i(\sigma )`$
this implies that all โmomentโ chemical potentials, $`\mu _if/\rho _i`$, are likewise equal among phases. The second requirement for phase coexistence is that the pressures or osmotic pressures , $`\mathrm{\Pi }`$, of all phases must also be equal. But from
$`\mathrm{\Pi }=f{\displaystyle ๐\sigma \mu (\sigma )\rho (\sigma )}=f{\displaystyle \underset{i}{}}\mu _i\rho _i`$
one sees that this again involves only the moment densities $`\rho _i`$ and their chemical potentials $`\mu _i`$. Finally, the moment densities also obey the โlever ruleโ, as follows. Let the overall density distribution of a system of volume $`V`$ be $`\rho ^{(0)}(\sigma )`$; we call this the โparentโ distribution. If (after a lowering of temperature, for example) this parent has split into $`p`$ coexisting โdaughterโ phases with $`\sigma `$-distributions $`\rho ^{(\alpha )}(\sigma )`$, each occupying a fraction $`v^{(\alpha )}`$ of the total volume ($`\alpha =1\mathrm{}p`$), then particle conservation implies the usual lever rule (or material balance) among species: $`_\alpha v^{(\alpha )}\rho ^{(\alpha )}(\sigma )=\rho ^{(0)}(\sigma ),\sigma `$. Multiplying this by a weight function $`w_i(\sigma )`$ and integrating over $`\sigma `$ shows that the lever rule also holds for the moment densities:
$$\underset{\alpha =1}{\overset{p}{}}v^{(\alpha )}\rho _i^{(\alpha )}=\rho _i^{(0)}$$
(2)
These results express the fact that any linear combination of conserved densities (a generalized moment density) is itself a conserved density in thermodynamics. We have shown, therefore, that if the free energy of the system depends only on $`K`$ moment densities $`\rho _1\mathrm{}\rho _K`$ we can view these as the densities of $`K`$ โquasi-speciesโ of particles, and construct the phase diagram via the usual construction of tangencies and the lever rule. Formally this has reduced the problem to finite dimensionality, although this is trivial here because $`f`$, by assumption, has no dependence on any variables other than the $`\rho _i`$ ($`i=1\mathrm{}K`$).
Of course, it is uncommon for the free energy $`f`$ to obey (1). In particular, the entropy of an ideal mixture (or, for polymers, the Flory-Huggins entropy term) is definitely not of this form. On the other hand, in very many thermodynamic (especially mean field) models the excess (i.e., non-ideal) part of the free energy does have the simple form (1). In other words, if we decompose the free energy as (setting $`k_\mathrm{B}=1`$)
$$f=Ts_{\mathrm{id}}+\stackrel{~}{f},s_{\mathrm{id}}=๐\sigma \rho (\sigma )\left[\mathrm{ln}\rho (\sigma )1\right]$$
(3)
then the excess free energy $`\stackrel{~}{f}`$ is a function of some moment densities only:
$$\stackrel{~}{f}=\stackrel{~}{f}(\rho _i),i=1\mathrm{}K$$
(4)
Examples of models of this kind include polydisperse hard spheres (within the generalization by Salacuse and Stell of the BMCSL equation of state ), polydisperse homo- and copolymers , and van der Waals fluids with factorized interaction parameters . With the exception of a brief discussion in Sec. VI, this paper concerns models with free energies of the form (3,4), which we call โtruncatableโ. (This terminology emphasizes that the number $`K`$ of moment densities appearing in the excess free energy of truncatable models is finite, while for a non-truncatable model the excess free energy depends on all details of $`\rho (\sigma )`$, corresponding to an infinite number of moment densities.) In what follows, we regard each model free energy as given, and do not discuss the issue of how good a description of the real system it offers, nor how or whether it can be derived from an underlying microscopic Hamiltonian. Whenever we refer to โexactโ results, we mean the exact thermodynamics of such a model as specified by its free energy.
Different authors refer in different terms to $`s_{\mathrm{id}}`$ in (3). Throughout this paper will call the quantity
$`s_{\mathrm{id}}={\displaystyle ๐\sigma \rho (\sigma )\left[\mathrm{ln}\rho (\sigma )1\right]}`$
(which is, up to a factor of $`T`$, the ideal part of the free energy density) the โentropy of an ideal mixtureโ. By writing $`\rho (\sigma )=\rho n(\sigma )`$, where $`n(\sigma )`$ is the normalized distribution of $`\sigma `$, this can be decomposed as
$$s_{\mathrm{id}}=\rho (\mathrm{ln}\rho 1)\rho ๐\sigma n(\sigma )\mathrm{ln}n(\sigma )$$
(5)
The first term on the right hand side is the โentropy of an ideal gasโ, while the second term gives the โentropy of mixingโ. The prefactor $`\rho `$ reflects the fact that both are expressed per unit volume rather than per particle.
In principle, the entire phase equilibria for any truncatable system (obeying (4)) can be computed exactly by a finite algorithm. Specifically, the spinodal stability criterion involves a $`K`$-dimensional square matrix whereas calculation of $`p`$-phase equilibrium involves solution of $`(p1)(K+1)`$ coupled nonlinear equations (see Sec. III E). This method has certainly proved useful , but is cumbersome, particularly if one is interested mainly in cloud- and shadow-curves, rather than coexisting compositions deep within multi-phase regions . Various ways of simplifying the procedure exist , but there has previously not been a systematic alternative to the full computation. Note also that the nonlinear phase equilibrium equations permit no simple geometrical interpretation or qualitative insight akin to the familiar rules for constructing phase diagrams from the free energy surface of a finite mixture. This further motivates the search for a description of polydisperse phase equilibria in terms of a reduced free energy which depends on only a small number of density variables.
In previous work, the authors originally arrived independently at two definitions of a reduced free energy in terms of moment densities (moment free energy for short) . Though based on distinctly different principles, the two approaches led to very similar results. We explain this somewhat surprising fact in the present work; at the same time, we describe the two methods in more detail and explore the relationship between them. We also discuss issues related to practical applications and give a number of example results for simple model systems.
The first route to a moment free energy takes as its starting point the conventional form of the ideal part of the free energy, $`f_{\mathrm{id}}=T๐\sigma \rho (\sigma )[\mathrm{ln}\rho (\sigma )1]`$, as in (3). This can be thought as defining a (hyper-)surface, with the โhorizontalโ co-ordinate being $`\rho (\sigma )`$ and the โheightโ of the surface $`f_{\mathrm{id}}[\rho (\sigma )]`$. As explained in Sec. II A, this surface can then be projected geometrically onto one with only a finite set of horizontal co-ordinates: these are chosen to be the moment densities appearing in the excess free energy. Physically, this corresponds to minimizing the free energy with respect to all degrees of freedom in $`\rho (\sigma )`$ that do not affect these moment densities. We call this first route the โprojectionโ method.
The second approach rederives the entropy of mixing in the ideal part of the free energy in a form that depends explicitly only on the chosen moment densities. As described in Sec. II B, the expression that results is intractable in general because it still contains the full complexity of the problem. However, in situations where there are only infinitesimal amounts of all but one of the phases in the systems, the entropy of mixing can be evaluated in a closed form. Applying this functional form in regimes where it is not strictly valid (i.e., when phases of comparable volumes coexist) generates a moment free energy by this โcombinatorialโ method.
We show the equivalence of the two approaches in Sec. II C. There, we first demonstrate that the general form of the entropy of mixing obtained by the combinatorial method can be transformed to the standard expression $`๐\sigma n(\sigma )\mathrm{ln}n(\sigma )`$. Second, we show that the moment free energies arrived at by the two methods are in fact equal, with the projection method being slightly more generally applicable.
In Sec. III, we then discuss the properties of the moment free energy (as obtained from either the projection or the combinatorial method). By construction, it depends only on a (finite) number of moment densities, and so achieves the desired reduction in dimensionality. The moment densities can be treated as densities of โquasi-speciesโ of particles, and the standard procedures of the thermodynamics of finite mixtures can be applied to the moment free energy to calculate phase equilibria. This is only useful, however, if the results are faithful to the actual phase behavior of the underlying polydisperse system (as modeled by the given truncatable free energy). We show that this is so: In fact, the construction of our moment free energy is such that exact binodals (cloud-point and shadow curves), critical (and multi-critical) points and spinodals are obtained. Beyond the onset of phase separation, where coexisting phases occupy comparable volumes, the results are not exact, but can be refined arbitrarily by adding extra moment densities. This procedure is necessary also to ensure that, in regions of multi-phase coexistence, the correct number of phases is found. In Sec. IV, we discuss the practical implementation of our method, followed by a number of examples (Sec. V). Our results are summarized in Sec. VI, where we also outline perspectives for future work.
Note that throughout this work, we focus on the case of phase coexistence at fixed volume. However, as described in App. A, the formalism can easily be applied to scenarios where the (mechanical or osmotic) pressure is fixed instead.
## II Derivation of moment free energy
### A Projection method
The starting point for this method is the decomposition (3) for the free energy of truncatable systems (4):
$$f=T๐\sigma \rho (\sigma )\left[\mathrm{ln}\frac{\rho (\sigma )}{R(\sigma )}1\right]+\stackrel{~}{f}(\rho _i)$$
(6)
As explained in the previous section, truncatable here means that the excess free energy $`\stackrel{~}{f}`$ depends only on $`K`$ moment densities $`\rho _i`$. Note that, in the first (ideal) term of (6), we have included a dimensional factor $`R(\sigma )`$ inside the logarithm. This is equivalent to subtracting $`T๐\sigma \rho (\sigma )\mathrm{ln}R(\sigma )`$ from the free energy. Since this term is linear in densities, it has no effect on the exact thermodynamics; it contributes harmless additive constants to the chemical potentials $`\mu (\sigma )`$. However, in the projection route to a moment free energy, it will play a central role.
We now argue that the most important moment densities to treat correctly are those that actually appear in the excess free energy $`\stackrel{~}{f}(\rho _i)`$. Accordingly we divide the infinite-dimensional space of density distributions into two complementary subspaces: a โmoment subspaceโ, which contains all the degrees of freedom of $`\rho (\sigma )`$ that contribute to the moment densities $`\rho _i`$, and a โtransverse subspaceโ which contains all remaining degrees of freedom (those that can be varied without affecting the chosen moment densities $`\rho _i`$. Physically, it is reasonable to expect that these โleftoverโ degrees of freedom play a subsidiary role in the phase equilibria of the system, a view justified a posteriori below. Accordingly, we now allow violations of the lever rule, so long as these occur solely in the transverse space. This means that the phase splits that we calculate using this approach obey the lever rule for the moment densities, but are allowed to violate it in other details of the density distribution $`\rho (\sigma )`$ (see Fig. 10 in Sec. V B 1 for an example). These โtransverseโ degrees of freedom, instead of obeying the strict particle conservation laws, are chosen so as to minimize the free energy: they are treated as โannealedโ. If, as assumed above, $`\stackrel{~}{f}=\stackrel{~}{f}(\rho _i)`$ only depends on the moment densities retained, this amounts to maximizing the ideal mixture entropy ($`s_{\mathrm{id}}`$) in (6), while holding fixed the values of the moment densities $`\rho _i`$. Note that we are allowed, if we wish, to include among the retained densities โredundantโ moments on which $`\stackrel{~}{f}`$ has a null dependence. We will have occasion to do this later on.
At this point, the factor $`R(\sigma )`$ in (6), which is immaterial if all conservation laws are strictly obeyed, becomes central. Indeed, maximizing the entropy over all distributions $`\rho (\sigma )`$ at fixed moment densities $`\rho _i`$ yields
$$\rho (\sigma )=R(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right)$$
(7)
where the Lagrange multipliers $`\lambda _i`$ are chosen to give the desired moment densities
$$\rho _i=๐\sigma w_i(\sigma )R(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right)$$
(8)
The corresponding minimum value of $`f`$ then defines our projected free energy as a function of the moment densities $`\rho _i`$:
$$f_{\mathrm{pr}}(\rho _i)=Ts_{\mathrm{pr}}(\rho _i)+\stackrel{~}{f}(\rho _i),s_{\mathrm{pr}}=\rho _0\underset{i}{}\lambda _i\rho _i$$
(9)
Here $`s_{\mathrm{pr}}`$ is the projected entropy of an ideal mixture. The first term appearing in it, $`\rho _0=๐\sigma \rho (\sigma )`$, is the โzeroth momentโ, which is identical to the overall particle density $`\rho `$ defined previously. If this is among the moment densities used for the projection (or more generally, if it is a linear combination of them), then the term $`T\rho _0`$ is simply a linear contribution to the projected free energy $`f_{\mathrm{pr}}(\rho _i)`$ and can be dropped because it does not affect phase equilibrium calculations. Otherwise, $`\rho _0`$ needs to be expressed โ via the $`\lambda _i`$ โ as a function of the $`\rho _i`$ and its contribution cannot be ignored. We will see an example of this in Sec. V.
The projection method yields a free energy $`f_{\mathrm{pr}}(\rho _i)`$ which only depends on the chosen set of moment densities. These can now be viewed as densities of โquasi-speciesโ of particles, and a finite-dimensional phase diagram can be constructed from $`f_{\mathrm{pr}}`$ according to the usual rules, ignoring the underlying polydisperse nature of the system. Obviously, though, the results now depend on $`R(\sigma )`$ which is formally a โprior distributionโ for the entropy maximization. To understand its role, we recall that the projected free energy $`f_{\mathrm{pr}}(\rho _i)`$ was constructed as the minimum of $`f[\rho (\sigma )]`$ at fixed $`\rho _i`$; that is, $`f_{\mathrm{pr}}`$ is the lower envelope of the projection of $`f`$ onto the moment subspace. Crucially, the shape of this envelope depends on how, by choosing a particular prior distribution $`R(\sigma )`$, we โtiltโ the infinite-dimensional free energy surface before projecting it. This geometrical point of view is illustrated, for a mixture of only two species, in Fig. 1.
To understand the effect of the prior $`R(\sigma )`$ physically, we note that the projected free energy is simply the free energy of phases of a system in which the density distributions $`\rho (\sigma )`$ are of the form (7). The prior $`R(\sigma )`$ determines which distributions lie within this โmaximum entropy familyโ (or โfamilyโ for short), and it is the properties of phases with these distributions that the projected free energy represents. Typically, one is interested in a system where a fixed overall โparentโ (or โfeedโ) distribution $`\rho ^{(0)}(\sigma )`$ becomes subject to separation into various phases. In such circumstances, we should generally choose this parent distribution as our prior, $`R(\sigma )=\rho ^{(0)}(\sigma )`$, thereby guaranteeing that it is contained within the family (7). Having done this, we note that the projection procedure will be exactly valid, to whatever extent the density distributions actually arising in the various coexisting phases of the system under study are members of the corresponding family
$$\rho (\sigma )=\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right).$$
(10)
In fact, the condition just described holds whenever all but one of a set of coexisting phases are of infinitesimal volume compared to the majority phase. This is because the density distribution, $`\rho ^{(0)}(\sigma )`$, of the majority phase is negligibly perturbed, whereas that in each minority phase differs from this by a Gibbs-Boltzmann factor, of exactly the form required for (10); we show this formally in Sec. III. Accordingly, our projection method yields exact cloud point and shadow curves. By the same argument, critical points (which in fact lie at the intersection of these two curves) are exactly determined; the same is true for tricritical and higher order critical points. Finally, spinodals are also found exactly. We defer explicit proofs of these statements to Sec. III.
The projection method does, however, give only approximate results for coexistences involving finite amounts of different phases. This is because linear combinations of different density distributions from the family (10), corresponding to two (or more) phases arising from the same parent $`\rho ^{(0)}(\sigma )`$, do not necessarily add to recover the parent distribution itself:
$$\underset{\alpha }{}v^{(\alpha )}\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _i^{(\alpha )}w_i(\sigma )\right)\rho ^{(0)}(\sigma )$$
(11)
unless all except one of the $`v^{(\alpha )}`$ are infinitesimal. Moreover, according to Gibbsโ phase rule, a projected free energy depending on $`n`$ moment densities will not normally predict more than $`n+1`$ coexisting phases, whereas a polydisperse system can in principle separate into an arbitrary number of phases. We explain in Sec. III how both of these shortcomings can be overcome by systematically including extra moment densities within the projection procedure. How quickly convergence to the exact results occurs depends on the choice of weight functions for the extra moment densities; we discuss this point further in the context of the examples of Sec. V.
### B Combinatorial method
Now we turn to the second method of constructing a moment free energy. As before, we recognize from the outset that the physics of the problem is contained in the excess free energy and therefore the most important variables are the moment densities which feature in this. On the other hand, the ideal part of the free energy (essentially the entropy of mixing, up to the trivial ideal gas term) is a book-keeping device that accounts for the number of ways of partitioning a priori a given size distribution between two or more coexisting phases. (To focus the discussion, we identify $`\sigma `$ as particle size in this section; the arguments do of course remain valid for a general polydispersity parameter $`\sigma `$.) The entropy of mixing has its origin in $`1/N!`$ factors in the partition function (where $`N`$ is a particle number) which are usually derived from the classical limit of quantum statistics. The origin of the $`1/N!`$ factors in classical statistical mechanics has been the subject of debate ever since Gibbs . The problem is also connected with the question of the extensivity of entropy, and the Gibbs paradox. Below we present a completely classical derivation of the entropy of mixing which has the additional benefit of indicating the appropriate generalisation for moment densities. To start with, follow Gibbs and define a non-extensive free energy $`F^{}`$ as a configuration space ($`\mathrm{\Gamma }`$) integral
$$e^{F^{}/T}=๐\mathrm{\Gamma }e^{H/T}.$$
(12)
No $`1/N!`$ appears in this as all particles are distinguishable in classical statistical mechanics.
Now consider two phases in coexistence as one joint system. Assume they occupy volumes $`V^{(1)}`$ and $`V^{(2)}`$, respectively, and contain $`N^{(1)}`$ and $`N^{(2)}`$ particles. Following literally the prescription in eq. (12), the free energy of the joint system of $`N^{(1)}+N^{(2)}=N`$ particles is found from
$$e^{F^{}/T}=\underset{\mathrm{prtns}}{}e^{(F_{}^{}{}_{}{}^{(1)}+F_{}^{}{}_{}{}^{(2)})/T}$$
(13)
where the configuration space integral has been done in two parts. Firstly, for each way of partitioning the particles between the two phases, the individual configuration space integrals give a product of the individual partition functions. Secondly, and crucially, one must sum over the $`N!/N^{(1)}!N^{(2)}!`$ ways of partitioning the particles between phases. Now define an extensive free energy by reinserting the $`1/N!`$ as though the particles were indistinguishable:
$$e^{F_{\mathrm{in}}/T}=\frac{1}{N!}๐\mathrm{\Gamma }e^{H/T},$$
(14)
Eq. (13) can then be written as
$$e^{F_{\mathrm{in}}/T}=e^{(F_{\mathrm{in}}^{(1)}+F_{\mathrm{in}}^{(2)})/T}_{\mathrm{prtns}}$$
(15)
where the average is taken over all partitions, with equal a priori probabilities. This result is the cornerstone of the combinatorial method. If we separate $`F_{\mathrm{in}}`$ into the entropy of an ideal gas and the remaining non-ideal (excess) parts, it is written as
$$F_{\mathrm{in}}=NT[\mathrm{ln}(N/V)1]+\stackrel{~}{F}$$
(16)
with $`\stackrel{~}{F}`$ being the excess free energy. Note that $`F_{\mathrm{in}}`$ does not yet contain an entropy of mixing term because it treats the particles as though they are indistinguishable. The entropy of mixing will reappear when we come to do the average over partitions in (15). Note also that conventional thermodynamics follows from (15) if all particles are identical as far as their mutual interactions are concerned. There is then no entropy of mixing to consider, so that $`F_{\mathrm{in}}`$ is just the conventional free energy $`F`$; and since the average over partitions is trivial one has $`F=F^{(1)}+F^{(2)}`$. Similarly, as we show in Sec. II C, one can recover from (15) the conventional form of the entropy of mixing as given in (5).
We now show how the average over partitions in (15) results in an expression for the entropy of mixing which depends explicitly on moment variables. The key is to note that, by our assumption that we are dealing with a truncatable model, the excess free energy $`\stackrel{~}{F}=V\stackrel{~}{f}(\rho _i)`$ depends only on a limited number of moment densities $`\rho _i`$ (as well as volume $`V`$, temperature $`T`$, and other state variables which we suppress below). If the density $`\rho _0`$ is among the moment densities $`\rho _i`$ โ which we assume throughout this section โ then specifying $`V`$ and the $`\rho _i`$ is equivalent to specifying $`N`$, $`V`$ and the normalized moments $`m_i=\rho _i/\rho _0`$ ($`i>0`$). So we can think of $`\stackrel{~}{F}`$ as a function of $`N`$, $`V`$ and the $`m_i`$. From (16), $`F_{\mathrm{in}}`$ then depends on the same set of variables. The choice of the $`m_i`$, which are moments of the normalized particle size distribution $`n(\sigma )`$, as independent variables is natural in the present context because we are considering the particle number $`N`$ and volume $`V`$ of the coexisting phases to be fixed.
To avoid notational complexity, and without loss of generality, we now specialize to the case where there is only one normalized moment $`m_i`$, a generalized mean size
$`m={\displaystyle ๐\sigma w(\sigma )n(\sigma )}`$
We shall also write $`\rho _m=\rho m=๐\sigma w(\sigma )\rho (\sigma )`$ for the corresponding moment density, and $`m^{(1)}`$ and $`m^{(2)}`$ for the values of the generalized mean size in the first and second phase. The overall (โparentโ) size distribution is $`n^{(0)}(\sigma )`$ with mean $`m^{(0)}`$. In each of the two phases, in which $`N`$ and $`V`$ are held fixed, $`F_{\mathrm{in}}`$ only depends on $`m`$, and so the average over partitions becomes
$$e^{(F_{\mathrm{in}}^{(1)}+F_{\mathrm{in}}^{(2)})/T}_{\mathrm{prtns}}๐m^{(1)}๐ซ(m^{(1)})e^{(F_{\mathrm{in}}^{(1)}+F_{\mathrm{in}}^{(2)})/T}$$
(17)
Here $`๐ซ(m^{(1)})`$ is the probability distribution for the generalized mean size in the first phase, taken over partitions with fixed $`N^{(1)}`$ and $`N^{(2)}`$, with equal a priori probabilities. Note that given $`m^{(1)}`$, $`m^{(2)}`$ is fixed in the second phase by the moment equivalent of particle conservation: $`N^{(1)}m^{(1)}+N^{(2)}m^{(2)}=Nm^{(0)}`$. The integral in (17) can be replaced by the maximum of the integrand in the thermodynamic limit, because $`\mathrm{ln}๐ซ(m^{(1)})`$ is an extensive quantity. Introducing a Lagrange multiplier $`\mu _m`$ for the above moment constraint then shows that the quantity $`\rho _m`$ has the same status as the density $`\rho \rho _0`$ itself: both are thermodynamic density variables. This reinforces the discussion in the introduction, where we showed that moment densities can be regarded as densities of โquasi-speciesโ of particles.
In what follows, we only need to refer to $`m^{(1)}`$ (not $`m^{(2)}`$) and therefore drop the superscript for brevity. Since $`\mathrm{ln}๐ซ(m)`$ is extensive, we can write $`\mathrm{ln}๐ซ(m)=Ns(m)`$ where $`s`$ is the entropy of mixing per particle expressed as a function of the moment $`m`$ (or, more generally, of the full set of normalized moments). The total free energy of the system then takes the form
$$F_{\mathrm{in}}=\underset{m}{\mathrm{min}}F_{\mathrm{in}}^{(1)}+F_{\mathrm{in}}^{(2)}NTs(m),$$
(18)
and we need to calculate the entropy of mixing $`s(m)`$. We recognize that this quantity depends not only on the generalized mean size, but also on $`x=N^{(1)}/N`$, the fraction of particles in the first phase. A formal result for $`s(m)`$ can be obtained by first calculating the joint probability distribution $`๐ซ(xm,x)`$ of $`xm`$ and $`x`$. From this, we can find $`๐ซ(m)`$ according to
$$๐ซ(m)=\frac{๐ซ(xm,x)}{๐m^{}๐ซ(xm^{},x)}=\frac{๐ซ(xm,x)}{๐ซ(x)/x}$$
(19)
Writing the log probabilities in (19) as $`\mathrm{ln}๐ซ(xm,x)=Ns(xm,x)`$ and $`\mathrm{ln}๐ซ(x)=Ns(x)`$, we then have in the thermodynamic limit
$$s(m)=s(xm,x)s(x)$$
(20)
The method for calculating $`๐ซ(xm,x)`$ proceeds by introducing an indicator function $`ฯต_i`$ for each particle, deemed to be $`1`$ if the particle is in the first phase, and $`0`$ if the particle is in the second . Then $`Nx=_{i=1}^Nฯต_i`$ and $`Nxm=_{i=1}^Nฯต_iw(\sigma _i)`$. We now write the moment generating function for $`๐ซ(xm,m)`$ as follows
$`{\displaystyle d(xm)๐x๐ซ(xm,x)\mathrm{exp}[N(\theta x+\lambda xm)]}`$ $`=\mathrm{exp}\left[\theta _{i=1}^Nฯต_i+\lambda _{i=1}^Nฯต_iw(\sigma _i)\right]_{\mathrm{prtns}}`$ (23)
$`={\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{exp}\{ฯต_i[\theta +\lambda w(\sigma _i)]\}_{\mathrm{prtns}}=2^N{\displaystyle \underset{i=1}{\overset{N}{}}}\{1+\mathrm{exp}[\theta +\lambda w(\sigma _i)]\}`$
$`=2^N\mathrm{exp}\left({\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{ln}\{1+\mathrm{exp}[\theta +\lambda w(\sigma _i)]\}\right).`$
In the second equality we have used the fact that the $`ฯต_i`$ are independent discrete random variables taking the values $`0`$ and $`1`$ with a priori equal probabilities. Taking logarithms of (23) and dividing by $`N`$, we obtain on the right hand side an average over $`i=1\mathrm{}N`$ of a function of $`\sigma _i`$. Since the $`\sigma _i`$ are drawn independently from $`n^{(0)}(\sigma )`$, the law of large numbers guarantees that this average tends to the corresponding average over $`n^{(0)}(\sigma )`$ in the limit $`N\mathrm{}`$ (compare the discussion in Ref. ). We therefore have the final result
$$\frac{1}{N}\mathrm{ln}d(xm)๐x\mathrm{exp}\{N[s(xm,x)+\theta x+\lambda xm]\}=๐\sigma n^{(0)}(\sigma )\mathrm{ln}\{1+\mathrm{exp}[\theta +\lambda w(\sigma )]\}$$
(24)
where we have dropped an unimportant additive constant ($`\mathrm{ln}2`$) and have replaced $`๐ซ(xm,x)`$ by $`\mathrm{exp}[Ns(xm,x)]`$.
Now, the left hand side can be evaluated, by the steepest descent method, in the limit $`N\mathrm{}`$. Defining
$`\psi (\theta ,\lambda )={\displaystyle ๐\sigma n^{(0)}(\sigma )\mathrm{ln}\{1+\mathrm{exp}[\theta +\lambda w(\sigma )]\}}`$
we get $`s(xm,x)+\theta x+\lambda xm=\psi (\theta ,\lambda )`$ at the point where $`s/x+\theta =s/(xm)+\lambda =0`$. We recognize that the relationship between $`s(x,xm)`$ and $`\psi (\theta ,\lambda )`$ is a double Legendre transform. Inverting this shows that
$$s(xm,x)=\psi (\theta ,\lambda )\theta x\lambda xm$$
(25)
where
$$x=\psi /\theta xm=\psi /\lambda $$
(26)
Eq. (25) is essentially the desired result; to obtain $`s(m)`$ from (20), only $`s(x)`$ remains to be determined. It is obtained from $`๐ซ(x)`$ via $`\mathrm{exp}[Ns(x)]=๐ซ(x)=d(xm)๐ซ(xm,x)`$ which by inspection of the moment generating function is seen to correspond to the point $`\lambda =0`$. At this point, $`\psi =\mathrm{ln}(1+e^\theta )`$ and $`x=\psi /\theta =e^\theta /(1+e^\theta )`$. We find that $`\theta =\mathrm{ln}[x/(1x)]`$ and after some manipulation,
$$s(x)=\psi x\theta =x\mathrm{ln}x(1x)\mathrm{ln}(1x)$$
(27)
This is recognized as the standard entropy of mixing that is lost when a total of $`N`$ particles is partitioned into $`N^{(1)}=xN`$ and $`N^{(2)}=(1x)N`$ particles in two phases. The simple form of the result shows that the calculation leading to (25) is essentially a generalization of the Stirling approximation.
Inserting (25) and (27) into (20), we finally have the desired result for $`s(m)`$,
$$s(m)=\psi (\theta ,\lambda )\theta x\lambda xm+x\mathrm{ln}x+(1x)\mathrm{ln}(1x)$$
(28)
Unfortunately this is mainly formal because neither the integral defining $`\psi `$ nor the Legendre transform are likely to be tractable. However, we show in Sec. II C that (28) is equivalent to the more conventional form of the entropy of mixing as given by the second term in in eq. (5). From a conceptual point of view, it should be noted that the conventional form is normally derived by โbinningโ the distribution of particle sizes $`\sigma `$ and taking the number of bins to infinity after the thermodynamic limit $`N\mathrm{}`$ has been taken (see e.g. ). In our above derivation, on the other hand, we have assumed that even for finite $`N`$ all particles have different sizes $`\sigma _i`$, drawn randomly from $`n^{(0)}(\sigma )`$; the โpolydisperse limitโ is thus taken simultaneously with the thermodynamic limit. The relation between these two approaches โ which lead to the same results โ has been discussed in detail by Salacuse . Note that the first limit is physically more plausible for many homopolymer systems (where there may only be thousands or millions of species, with many particles of each) whereas the second limit is more natural for colloidal materials (and also some random copolymers) in which no two particles present are exactly alike, even in a sample of macroscopic size.
Although the full result (28) is intractable, progress can be made for $`x0`$, when the number of particles in one phase is much smaller than in the other. The limit $`x0`$ implies $`\theta \mathrm{}`$ and hence $`\psi (\theta ,\lambda )\mathrm{exp}[\theta +h(\lambda )]`$ where $`h=\mathrm{ln}๐\sigma n^{(0)}(\sigma )e^{\lambda w(\sigma )}`$ is a generalized cumulant generating function for $`n^{(0)}(\sigma )`$. From this we derive $`x=\psi /\theta =\psi `$, giving $`\theta =\mathrm{ln}xh`$. The generalized mean size is given by $`xm=\psi /\lambda =\psi h/\lambda `$ and hence $`m=h/\lambda `$. Inserting these results into (25) then shows that
$`s(xm,x)=xx\mathrm{ln}x+x(h\lambda m),m={\displaystyle \frac{h}{\lambda }}`$
and hence from (28)
$$s(m)=x(h\lambda m)+๐ช(x^2).$$
(29)
Noting that $`xN=N^{(1)}`$ is the number of particles in the small phase, the total free energy (18) can then be written as
$$F_{\mathrm{in}}=F_{\mathrm{in}}^{(2)}+F_{\mathrm{in}}^{(1)}N^{(1)}T(h\lambda m)$$
(30)
The term $`(h\lambda m)`$, being multiplied by $`N^{(1)}T`$, can now be interpreted as the entropy of mixing *per particle in the small phase*. It arises from the deviation of the (generalized) mean size $`mm^{(1)}`$ in the small phase from the mean size $`m^{(0)}`$ of the parent, and is given by the Legendre transform of the generalized cumulant generating function:
$$s_{\mathrm{mix}}(m)=h\lambda m,\text{where}m=\frac{h}{\lambda },h=\mathrm{ln}๐\sigma n^{(0)}(\sigma )e^{\lambda w(\sigma )}.$$
(31)
Let us now examine the relation between $`s_{\mathrm{mix}}(m)`$, thereby defined, and $`s(m)`$ introduced previously. By construction, $`Ns(m)`$ is the entropy of mixing of the system as a whole; both phases contribute to this. The result (29), which can also be written as
$$Ns(m)=N^{(1)}s_{\mathrm{mix}}(m)$$
(32)
appears to contradict this, seeing as it does not contain a term proportional to $`N^{(2)}`$. The resolution of this paradox comes from the neglected $`๐ช(x^2)`$ terms in (29). In fact, using the single-phase entropy of mixing defined in eq. (31) โ and reinstating on the r.h.s. the superscript on $`mm^{(1)}`$ โ we can write (32) in the more symmetric form
$`Ns(m)=N^{(1)}s_{\mathrm{mix}}(m^{(1)})+N^{(2)}s_{\mathrm{mix}}(m^{(2)})`$
This is still correct to leading (linear) order in $`x`$ because the added term is $`๐ช(x^2)`$. \[To see this, use the fact that $`m^{(2)}=m^{(0)}xm^{(1)}/(1x)`$ and hence $`m^{(2)}m^{(0)}=๐ช(x)`$. From (31) it can then be deduced that $`s_{\mathrm{mix}}(m^{(2)})(m^{(2)}m^{(0)})^2=๐ช(x^2)`$.\] Similarly, we can rewrite the total free energy (30) as
$$F_{\mathrm{in}}=[F_{\mathrm{in}}^{(1)}N^{(1)}Ts_{\mathrm{mix}}(m^{(2)})]+[F_{\mathrm{in}}^{(2)}N^{(2)}Ts_{\mathrm{mix}}(m^{(2)})]$$
(33)
This shows that in the limit where one of the two phases is much larger than the other one, the free energy of each of the phases โ now including the entropy of mixing โ is given by
$$F=F_{\mathrm{in}}NTs_{\mathrm{mix}}(m)=NT\left[\mathrm{ln}(N/V)+1+s_{\mathrm{mix}}(m)\right]+\stackrel{~}{F}$$
(34)
This expression, which only depends on the particle size distribution through the generalized mean size $`m`$, is our desired moment free energy.
The Legendre transform result (31) for the (single-phase) entropy of mixing is appealingly simple; we will illustrate its application in Sec. V A, using polydisperse Flory-Huggins theory as an example. As discussed briefly in App. B, eq. (31) also establishes an interesting connection to large deviation theory. However, the most important aspect of the result (31,34) is that โ as we have shown โ it gives exact results for the limiting case $`N^{(1)}N^{(2)}`$ (that is, $`x0`$). This includes two important classes of problems, that are also handled exactly by the projection method (for essentially the same reason). The first is the determination of spinodal curves and critical points. Intuitively these are exact because they are related to the stability of the system with respect to small variations in its density or composition, which can be probed by allowing fluctuations to take place in a vanishingly small subregion. The second concerns the cloud point and shadow curves. Again intuitively, these are exact because by definition only an infinitesimal amount of a second phase has appeared. Formal proofs of these statements are given in Sec. III.
If the result (31,33,34) is applied in the regime where it is no longer strictly valid, i.e., where the two coexisting phases contain comparable number of particles (or, equivalently, occupy comparable volumes), then approximate two-phase coexistences can be calculated. It is also straightforward to show that the analog of eqs. (31,33,34) holds in the case where two or more small phases coexist with a much larger phase; application in the regime of comparable phase volumes then provides approximate results for multi-phase coexistence.
To end this section, let us state in full the analogue of (31,34) for the case of several moment densities, restoring the notation used in the previous sections. The square bracket on the r.h.s. of (34) is the moment expression for the entropy of an ideal mixture. If, as in Sec. II A, we measure this entropy per unit volume (rather than per particle, as previously in the current Section) and generalize to several moment densities, we find by the combinatorial approach the following moment free energy:
$$f_{\mathrm{comb}}=Ts_{\mathrm{comb}}+\stackrel{~}{f},s_{\mathrm{comb}}=\rho _0(\mathrm{ln}\rho _01)+\rho _0\left(h\underset{i0}{}\lambda _im_i\right),$$
(35)
with
$$h=\mathrm{ln}๐\sigma n^{(0)}(\sigma )\mathrm{exp}\left(\underset{i0}{}\lambda _iw_i(\sigma )\right)$$
(36)
Here $`\rho _0=N/V`$ is the particle density as before.
### C Relation between the two methods
Although eq. (35) looks somewhat different from the corresponding result (9) for $`f_{\mathrm{pr}}`$ obtained by the projection method, the two methods are, mathematically, almost equivalent, as we now show. (For brevity we return to the case of a single moment $`m`$ and continue to refer to the polydisperse feature $`\sigma `$ as โsizeโ.) First, consider the exact expression for the entropy of mixing (28) derived within the combinatorial method. This should be equivalent to the conventional result used as the starting point (5) in the projection method. To see this, write the Legendre transform conditions (26) out explicitly:
$$x=๐\sigma n^{(0)}(\sigma )\frac{\mathrm{exp}[\theta +\lambda w(\sigma )]}{1+\mathrm{exp}[\theta +\lambda w(\sigma )]},xm^{(1)}=๐\sigma w(\sigma )n^{(0)}(\sigma )\frac{\mathrm{exp}[\theta +\lambda w(\sigma )]}{1+\mathrm{exp}[\theta +\lambda w(\sigma )]}$$
(37)
Here we have reinstated the superscript to show that $`m^{(1)}`$ is the value in phase one of the (generalized) moment $`m`$:
$`m^{(1)}={\displaystyle ๐\sigma w(\sigma )n^{(1)}(\sigma )}`$
Comparing with (37), we can identify the (normalized) particle size distribution in that phase as
$`n^{(1)}(\sigma )={\displaystyle \frac{n^{(0)}(\sigma )}{x}}{\displaystyle \frac{\mathrm{exp}[\theta +\lambda w(\sigma )]}{1+\mathrm{exp}[\theta +\lambda w(\sigma )]}}`$
Particle conservation $`xn^{(1)}(\sigma )+(1x)n^{(2)}(\sigma )=n^{(0)}(\sigma )`$ implies a similar form for the size distribution in phase two:
$`n^{(2)}(\sigma )={\displaystyle \frac{n^{(0)}(\sigma )}{1x}}{\displaystyle \frac{1}{1+\mathrm{exp}[\theta +\lambda w(\sigma )]}}.`$
It then takes only a few lines of algebra to show that the entropy of mixing (28) can be written (for any $`x`$) as
$`s(m)`$ $`=`$ $`x{\displaystyle ๐\sigma n^{(1)}(\sigma )\mathrm{ln}\frac{n^{(1)}(\sigma )}{n^{(0)}(\sigma )}}(1x){\displaystyle ๐\sigma n^{(2)}(\sigma )\mathrm{ln}\frac{n^{(2)}(\sigma )}{n^{(0)}(\sigma )}}`$ (38)
$`=`$ $`x{\displaystyle ๐\sigma n^{(1)}(\sigma )\mathrm{ln}n^{(1)}(\sigma )}(1x){\displaystyle ๐\sigma n^{(2)}(\sigma )\mathrm{ln}n^{(2)}(\sigma )}+{\displaystyle ๐\sigma n^{(0)}(\sigma )\mathrm{ln}n^{(0)}(\sigma )}`$ (39)
The second equation, with the third (constant) term discarded, is the standard result (compare eqs. (5,6,A1)). However the first expression appears more natural, and is retained, within the combinatorial derivation; it embodies the intuitively reasonable prescription that entropy is best measured relative to the parent distribution $`n^{(0)}(\sigma )`$. This was also an essential ingredient of the projection method, as described in Sec. II A, where the โpriorโ in the entropy expression was likewise identified with the parent. Retaining the parent as prior also avoids subtleties with the definition of the integrals in (39) in the case where the phases contain monodisperse components, corresponding to $`\delta `$-peaks in the density distributions .
Having established that the rigorous starting points of the combinatorial and projection method are closely related, we now show that the subsequent approximations also lead to essentially the same results. The relevant approximations are use of (35) for $`f_{\mathrm{comb}}`$ (which adopts the small $`x`$ form for $`s_{\mathrm{mix}}`$) in the combinatorial case and use of (9) for $`f_{\mathrm{pr}}`$ (which minimizes over transverse degrees of freedom) in the projection case. First note that the density $`\rho _0`$ of phases within the family (10) is given by
$`\rho _0={\displaystyle ๐\sigma \rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right)}`$
Comparing this with (36), and using the fact that $`n^{(0)}(\sigma )=\rho ^{(0)}(\sigma )/\rho _0^{(0)}`$, one sees that
$`\rho _0={\displaystyle ๐\sigma \rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right)}=\rho _0^{(0)}e^{\lambda _0+h}`$
where the sum over $`i`$ now includes $`i=0`$. Solving for $`h`$, one has $`h=\lambda _0+\mathrm{ln}(\rho _0/\rho _0^{(0)})`$. The ideal mixture entropy derived by the combinatorial route, eq. (35), can thus be rewritten as
$`s_{\mathrm{comb}}=\rho _0(\mathrm{ln}\rho _01)+\rho _0\mathrm{ln}(\rho _0/\rho _0^{(0)})\lambda _0\rho _0{\displaystyle \underset{i0}{}}\lambda _i\rho _0m_i=\rho _0{\displaystyle \underset{i}{}}\lambda _i\rho _i\rho _0\mathrm{ln}\rho _0^{(0)}.`$
This is identical to the projected entropy $`s_{\mathrm{pr}}`$, eq. (9), except for the last term. But by construction, the combinatorial entropy assumes that $`\rho _0`$ โ the overall density โ is among the moment densities retained in the moment free energy. The difference $`s_{\mathrm{comb}}s_{\mathrm{pr}}=\rho _0\mathrm{ln}\rho _0^{(0)}`$ is then linear in this density, and the combinatorial and projection methods therefore predict exactly the same phase behavior.
In summary, we have shown that the projection and combinatorial methods for obtaining moment free energies give equivalent results. The only difference between the two approaches is that within the projection approach, one need not necessarily retain the zeroth moment, which is the overall density $`\rho _0=\rho `$, as one of the moment densities on which the moment free energy depends. If $`\rho _0`$ does not appear in the excess free energy, this reduces the minimum number of independent variables of the moment free energy by one (see Sec. V for an example).
## III Properties of the moment free energy
In the previous sections, we have derived by two different routes (namely (9) and (35)) our moment free energy for truncatable polydisperse systems. We now investigate the properties of this moment free energy, which we henceforward denote $`f_\mathrm{m}`$ and write in the form (9)
$$f_\mathrm{m}(\rho _i)=Ts_\mathrm{m}(\rho _i)+\stackrel{~}{f}(\rho _i),s_\mathrm{m}=\rho _0\underset{i}{}\lambda _i\rho _i;$$
(40)
here $`s_\mathrm{m}`$ is the โmoment entropyโ of an ideal mixture. In particular, we formally compare the phase behavior predicted from this free energy (by treating the $`\rho _i`$ as densities of โquasi-speciesโ of particles, and applying the usual tangency construction) with that obtained from the exact free energy of the underlying (truncatable) model
$$f[\rho (\sigma )]=T๐\sigma \rho (\sigma )\left[\mathrm{ln}\rho (\sigma )1\right]+\stackrel{~}{f}(\rho _i).$$
(41)
Let us first collect a few simple properties of the moment free energy (40) which will be useful later. Recall that (40) faithfully represents the free energy density of any phase with density distribution in the family
$$\rho (\sigma )=\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _iw_i(\sigma )\right)$$
(42)
where $`\rho ^{(0)}(\sigma )`$ is the density distribution of the parent. The moment densities $`\rho _i`$ are then related to the Lagrange multipliers $`\lambda _i`$ by
$$\rho _i=๐\sigma w_i(\sigma )\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{j}{}\lambda _jw_j(\sigma )\right)$$
(43)
If we regard $`\rho _0`$ as a function of the $`\lambda _i`$, then from (43) we have $`\rho _0/\lambda _i=\rho _i`$. Together with eq. (40), this implies that the moment entropy $`s_\mathrm{m}`$ has the structure of a Legendre transform . For the first derivatives of $`s_\mathrm{m}`$ with respect to the moment densities, this yields
$`{\displaystyle \frac{s_\mathrm{m}}{\rho _i}}=\lambda _i`$
while the matrix of second derivatives is the negative inverse of the matrix of โsecond-order moment densitiesโ $`\rho _{ij}`$ :
$$\frac{^2s_\mathrm{m}}{\rho _i\rho _j}=(๐ด^1)_{ij},(๐ด)_{ij}=\frac{^2\rho _0}{\lambda _i\lambda _j}=๐\sigma w_i(\sigma )w_j(\sigma )\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{k}{}\lambda _kw_k(\sigma )\right)\rho _{ij}.$$
(44)
The chemical potentials $`\mu _i`$ conjugate to the moment densities follow as
$$\mu _i=\frac{f_\mathrm{m}}{\rho _i}=T\lambda _i+\frac{\stackrel{~}{f}}{\rho _i}=T\lambda _i+\stackrel{~}{\mu }_i$$
(45)
and their derivatives w.r.t. the $`\rho _i`$, which give the curvature of the moment free energy, are
$$\frac{\mu _i}{\rho _j}=\frac{^2f_\mathrm{m}}{\rho _i\rho _j}=T(๐ด^1)_{ij}+\frac{^2\stackrel{~}{f}}{\rho _i\rho _j}.$$
(46)
The pressure, finally, is given by
$$\mathrm{\Pi }_\mathrm{m}=f_\mathrm{m}\underset{i}{}\mu _i\rho _i=T\rho _0+\stackrel{~}{f}\underset{i}{}\stackrel{~}{\mu }_i\rho _i.$$
(47)
Eqs. (45-47) are all calculated via the moment free energy. The three corresponding quantities obtained from the exact free energy (41) are, first, the chemical potentials conjugate to $`\rho (\sigma )`$:
$$\mu (\sigma )=\frac{\delta f}{\delta \rho (\sigma )}=T\mathrm{ln}\rho (\sigma )+\underset{i}{}\stackrel{~}{\mu }_iw_i(\sigma )$$
(48)
second, their derivatives w.r.t. $`\rho (\sigma )`$:
$$\frac{\delta \mu (\sigma )}{\delta \rho (\sigma ^{})}=\frac{\delta ^2f}{\delta \rho (\sigma )\delta \rho (\sigma ^{})}=\frac{T\delta (\sigma \sigma ^{})}{\rho (\sigma )}+\underset{i,j}{}\frac{^2\stackrel{~}{f}}{\rho _i\rho _j}w_i(\sigma )w_j(\sigma ^{})$$
(49)
and, third, the resulting expression for the pressure:
$$\mathrm{\Pi }=f๐\sigma \mu (\sigma )\rho (\sigma )=T\rho _0+\stackrel{~}{f}\underset{i}{}\stackrel{~}{\mu }_i\rho _i.$$
(50)
The last of these is identical to the result (47) derived from the moment free energy.
We note one important consequence of the form of the exact chemical potentials (48) for truncatable models: If two phases $`\rho ^{(1)}(\sigma )`$ and $`\rho ^{(2)}(\sigma )`$ have the same chemical potentials $`\mu (\sigma )`$, then the ratio of their density distributions can be written as a Gibbs-Boltzmann factor:
$$\frac{\rho ^{(1)}(\sigma )}{\rho ^{(2)}(\sigma )}=\mathrm{exp}\left[\underset{i}{}\beta \left(\stackrel{~}{\mu }_i^{(2)}\stackrel{~}{\mu }_i^{(1)}\right)w_i(\sigma )\right]$$
(51)
This implies that if one of the density distributions is in the family (42), then so is the other. The same argument obviously applies if there are several phases with equal chemical potentials. Conversely, we have for the chemical potential difference between any two phases in the family (42)
$$\mathrm{\Delta }\mu (\sigma )=\underset{i}{}(T\mathrm{\Delta }\lambda _i+\mathrm{\Delta }\stackrel{~}{\mu }_i)w_i(\sigma )=\underset{i}{}\mathrm{\Delta }\mu _iw_i(\sigma ).$$
(52)
The chemical potentials $`\mu (\sigma )`$ of two such phases are therefore equal if and only if their moment chemical potentials $`\mu _i`$ are equal. Combining this with the fact that the pressure (47) derived from the moment free energy is exact (compare (50)), we conclude: Any set of (two or more) coexisting phases calculated from the moment free energy obeys the exact phase equilibrium conditions. That is, if they were brought into contact, they would genuinely coexist. (However, they will not necessarily obey the lever rule with respect to the parent $`\rho ^{(0)}(\sigma )`$ of the given family, unless all but one of them โ which then coincides with the parent โ are of vanishing volume; see (11).)
### A General criteria for spinodals and (multi-) critical points
We now demonstrate that, for any truncatable model, the moment free energy gives exact spinodals and (multi-) critical points. By this we mean that, using the moment free energy, the values of external control parameters โ such as temperature โ at which the parent phase $`\rho ^{(0)}(\sigma )`$ becomes unstable or critical can be exactly determined. Our argument treats spinodals and (multi-) critical points in a unified fashion, using the fact that all of them occur when the difference between phases with equal chemical potentials becomes infinitesimal. Because the truncatable models that we are considering are generally derived from mean-field approaches, we do not concern ourselves with critical point singularities, assuming instead that the free energy is a smooth function of all order parameters (the densities in the system). Likewise, we do not discuss the subtle question of how, beyond mean-field theory, free energies can actually be defined in spinodal and unstable regions .
Let us first recap the general criteria for spinodals and critical points in multi-component systems. In order to treat the critera derived from the exact and moment free energies simultaneously, we use the common notation $`๐`$ for the vector of densities specifying the system: for the exact free energy, the components of $`๐`$ are the values $`\rho (\sigma )`$; for the moment free energy, they are the reduced set $`\rho _i`$. We write the corresponding vector of chemical potentials as
$`๐(๐)=f(๐).`$
This notation emphasizes that the chemical potentials are functions of the densities. The criterion for a spinodal at the parent phase $`๐^{(0)}`$ is then that there is an incipient instability direction $`\delta ๐`$ along which the chemical potentials do not change:
$$(\delta ๐)๐(๐^{(0)})=(\delta ๐)f(๐^{(0)})=\mathrm{๐}.$$
(53)
As the second form of the criterion shows, an equivalent statement is that the curvature of the free energy along the direction $`\delta ๐`$ vanishes. Eq. (53) can also be written as
$$๐(๐^{(0)}+ฯต\delta ๐)๐(๐^{(0)})=๐ช(ฯต^2)$$
(54)
which is closely related to the critical point criterion that we describe next.
Near a critical point, the parent $`๐^{(0)}`$ coexists with another phase that is only slightly different; if, as we assume here, the free energy function is smooth, these two phases are separated โ in $`๐`$-space โ by a โhypothetical phaseโ which has the same chemical potentials but is (locally) thermodynamically unstable. (This is geometrically obvious even in high dimensions; between any two minima of $`f(๐)๐๐`$, at given $`๐`$, there must lie a maximum or a saddle point, which is the required unstable โphaseโ.) Now imagine connecting these three phases by a smooth curve in density space $`๐(ฯต)`$. At the critical point, all three phases collapse, and the variation of the chemical potential around $`๐(ฯต=0)=๐^{(0)}`$ must therefore obey
$`๐(๐(ฯต))๐(๐^{(0)})=๐ช(ฯต^3)`$
Similarly, if $`n`$ phases coexist, we can connect them and the $`n1`$ unstable phases in between them by a curve $`๐(ฯต)`$ . If we define an $`n`$-critical point as one where all these phases become simultaneously critical, we obtain the criterion
$$๐(๐(ฯต))๐(๐^{(0)})=๐ช(ฯต^{2n1}).$$
(55)
This formulation was proposed by Brannock ; the cases $`n=2`$ and $`n=3`$ correspond to ordinary critical and tricritical points, respectively. The spinodal criterion (54) is of the same form as (55) if one chooses the curve $`๐(ฯต)=๐^{(0)}+ฯต\delta ๐`$.
In summary, we have that the phase $`๐^{(0)}`$ is a spinodal or $`n`$-critical point if there is a curve $`๐(ฯต)`$ with $`๐(ฯต=0)=๐^{(0)}`$ such that
$$\mathrm{\Delta }๐๐(๐(ฯต))๐(๐^{(0)})=๐ช(ฯต^l).$$
(56)
where $`l=2`$ for a spinodal, and $`l=2n1`$ for an $`n`$-critical point. Brannock has shown that these criteria are equivalent to the determinant criteria introduced by Gibbs . The above forms are more useful for us because they avoid having to define infinite-dimensional determinants (as otherwise required to handle the exact free energy $`f(๐)`$ of a polydisperse system, whose argument $`๐`$ denotes an infinite number of components, even in the truncatable case). They also show the analogy with the standard criteria for single-species systems (where $`๐`$ has only a single component) more clearly.
We can now apply (56) to the exact free energy (41) and show that the resulting criteria are identical to those obtained from the moment free energy. First, note that the curve $`๐(ฯต)\rho (\sigma ;ฯต)`$ can always be chosen to lie within the family (42). This follows from the derivation of the criterion (56): The curve $`๐(ฯต)`$ is defined as passing through $`l`$ phases with equal chemical potentials, one of them being the parent $`๐^{(0)}`$. As shown in (51), all these phases are therefore within the family (42). But then (52) implies that the condition (56) that $`\mathrm{\Delta }\mu (\sigma )`$ must be zero to $`๐ช(ฯต^l)`$ is equivalent to the same requirement for the differences $`\mathrm{\Delta }\mu _i`$ in the moment chemical potential. This proves that the conditions for spinodals and (multi-) critical points derived from the exact and moment free energies are equivalent. Intuitively, one can understand this as follows: Having shown that the spinodal/critical point conditions can be formulated solely in terms of density distributions $`\rho (\sigma )`$ within the family (42), it is sufficient to know the free energy of those density distributions. This is exactly the moment free energy.
Finally, we note that the exactness of these stability and critical point conditions holds not just for the parent $`\rho ^{(0)}(\sigma )`$, but for all phases within the family (42). This is clear, because any such phase could itself be chosen as the parent without changing the family; the moment free energy would change only by irrelevant terms linear in the moment densities. In general, one will not necessarily be interested in the properties of such โsubstituteโ parents. As an exception, substitute parents that differ from the parent only by a change of the overall density are of interest: they lie on the dilution line of distributions from which cloud point and shadow curves are calculated. The dilution line is included in the family (42) if the overall density $`\rho _0`$ (with corresponding weight function $`w_0(\sigma )=1`$) is retained in the moment free energy; in the combinatorial derivation, this is automatically the case.
### B Spinodals
For completeness, we now give the explicit form of the spinodal criterion (53) for truncatable systems; see also for an equivalent derivation using the combinatorial approach. Using (46) and abbreviating the matrix of second derivatives of the excess free energy as $`\stackrel{\mathbf{~}}{๐ญ}`$, the spinodal condition (53) becomes
$$\left|\stackrel{\mathbf{~}}{๐ญ}+T๐ด^1\right|=0,(\stackrel{\mathbf{~}}{๐ญ}+T๐ด^1)๐น๐=\mathrm{๐}.$$
(57)
As before, the (nonzero) vector $`๐น๐`$ with components $`\delta \rho _i`$ gives the direction of the spinodal instability; the moment densities $`\rho _i`$ and $`\rho _{ij}`$ that appear in $`\stackrel{\mathbf{~}}{๐ญ}`$ and $`๐ด`$ are to be evaluated for the parent distribution $`\rho ^{(0)}(\sigma )`$ being studied. Note that (57) is valid for any $`\rho ^{(0)}(\sigma )`$; no specific assumptions about the parent were made in the derivation. We can thus simply drop the โ(0)โ superscript: for any phase with density distribution $`\rho (\sigma )`$, the point where (57) first becomes zero, as external control parameters are varied, locates a spinodal instability.
More convenient forms of (57) that avoid matrix inversions are obtained after multiplication by the second order moment matrix $`๐ด`$ (which is positive definite for linearly independent weight functions $`w_i(\sigma )`$ and therefore has nonzero determinant):
$$Y=\left|\mathrm{๐}+\beta ๐ด\stackrel{\mathbf{~}}{๐ญ}\right|=0,(\mathrm{๐}+\beta ๐ด\stackrel{\mathbf{~}}{๐ญ})๐น๐=\mathrm{๐}.$$
(58)
Here $`\beta =1/T`$ in the standard notation. From our general statements in Sec. III A, the spinodal criterion derived from the exact free energy (41) must be identical to this; this is shown explicitly in App. C. Note that the spinodal condition depends only on the (first-order) moment densities $`\rho _i`$ and the second-order moment densities $`\rho _{ij}`$ of the distribution $`\rho (\sigma )`$ (given by (43) and (44)); it is independent of any other of its properties. This simplification, which has been pointed out by a number of authors , is particularly useful for the case of power-law moments (defined by weight functions $`w_i(\sigma )=\sigma ^i`$): If the excess free energy only depends on the moments of order 0, 1$`\mathrm{}`$ $`K1`$ of the density distribution, the spinodal condition involves only $`2K1`$ moments \[up to order $`2(K1)`$\].
The general discussion in Sec. III A as well as the explicit calculation in App. C show that the spinodal instability direction lies within (or more precisely, is tangential to) the family (42) of density distributions. This fact has a simple geometrical interpretation: By construction, the family lies along a โvalleyโ of the free energy surface (compare Fig. 1). Away from the valley floor, any change in $`\rho (\sigma )`$ increases the free energy, corresponding to a positive curvature; the spinodal direction, for which the curvature vanishes, must therefore be along the valley floor.
### C Critical points
Next, we show the explicit form of the critical point criterion for truncatable systems. The general condition (56) for an ordinary critical point ($`n=2`$) was shown by Brannock to be equivalent to
$$(\delta ๐)f(๐^{(0)})=\mathrm{๐},(\delta ๐)^3f(๐^{(0)})=0$$
(59)
The first part of this is simply the spinodal criterion, as expected. To evaluate the second part for the moment free energy (40), we need the third derivative of $`s_\mathrm{m}`$ w.r.t. the moment densities $`\rho _i`$. Writing (44) as
$`{\displaystyle \underset{j}{}}\rho _{ij}{\displaystyle \frac{^2s_\mathrm{m}}{\rho _j\rho _k}}=\delta _{ik}`$
and differentiating w.r.t. one of the moment densities, one finds after a little algebra
$`{\displaystyle \frac{^3s_\mathrm{m}}{\rho _i\rho _j\rho _k}}={\displaystyle \underset{lmn}{}}(๐ด^1)_{il}(๐ด^1)_{jm}(๐ด^1)_{kn}\rho _{lmn}`$
with third-order moment densities $`\rho _{lmn}`$ defined in the obvious way. Thus, evaluating (59) for the moment free energy (40), we find that critical points have to obey
$$T\underset{ijk}{}v_iv_jv_k\rho _{ijk}+\delta \rho _i\delta \rho _j\delta \rho _k\frac{^3\stackrel{~}{f}}{\rho _i\rho _j\rho _k}=0,v_i=\underset{j}{}(๐ด^1)_{ij}\delta \rho _j$$
(60)
in addition to the spinodal condition (58). As before, the criterion has to be evaluated for the parent phase $`\rho ^{(0)}(\sigma )`$ under consideration; but because $`\rho ^{(0)}(\sigma )`$ can be chosen arbitrarily, it applies to all density distributions $`\rho (\sigma )`$.
As expected from the general discussion in Sec. III A, the criterion (60) can also be derived from the exact free energy; an alternative form involving the spinodal determinant $`Y`$ is given in App. D. Eq. (60) shows that the location of critical points depend only on the moment densities $`\rho _i`$, $`\rho _{ij}`$ and $`\rho _{ijk}`$ . For a system with an excess free energy depending only on power-law moments up to order $`K1`$, the critical point condition thus involves power-law moments of the parent only up to order $`3(K1)`$.
### D Onset of phase coexistence: Cloud point and shadow
So far in this section, we have shown that the moment free energy gives exact results for spinodals and (multi-) critical points. Now we consider the onset of phase coexistence, where (on varying the temperature, for example) a parent phase with density distribution $`\rho ^{(0)}(\sigma )`$ first starts to phase separate. As explained in Sec. I, this temperature together with the overall density $`\rho ^{(0)}\rho _0^{(0)}`$ of the parent defines a โcloud pointโ; the density of the incipient daughter phase gives the โshadowโ. If the parent begins to coexist with $`p`$ phases simultaneously ($`p=2`$ at a triple point, for example), there will be $`p`$ such shadows.
At the onset of phase coexistence, one of the coexisting phases is by definition the parent $`\rho ^{(0)}(\sigma )`$; the lever rule does not yet play any role because the daughters $`\rho ^{(\alpha )}(\sigma )`$ ($`\alpha =1\mathrm{}p`$) occupy an infinitesimal fraction of the total volume. It then follows from (51) that all daughters lie within the family (42) of density distributions. As shown in (52), the condition for equality of chemical potentials $`\mu (\sigma )`$ between any two of the coexisting phases (parent and daughters) then becomes
$`\mathrm{\Delta }\mu (\sigma )={\displaystyle \underset{i}{}}\mathrm{\Delta }\mu _iw_i(\sigma )`$
and is satisfied if and only if the moment chemical potentials $`\mu _i`$ are equal in all phases. Likewise, the exact and moment free energies give the same condition for equality of pressure $`\mathrm{\Pi }`$ in all phases, because they yield identical expressions (47,50) for $`\mathrm{\Pi }`$. In summary, we see that the conditions for the onset of phase coexistence are identical for the exact and moment free energies; the moment free energy therefore gives exact cloud points and shadows.
### E Phase coexistence beyond onset
As stated in Sec. II A, the moment free energy does not give exact results beyond the onset of phase coexistence, i.e., in the regime where the coexisting phases occupy comparable fractions of the total system volume. As shown in Sec. III A, the calculated phases will still be in exact thermal equilibrium; but the lever rule will now be violated for the โtransverseโ degrees of freedom of the density distributions. This is clear from (11): In general, no linear combination of distributions from this family can match the parent $`\rho ^{(0)}(\sigma )`$ exactly.
A more detailed understanding of the failure of the moment free energy beyond phase coexistence can be gained by comparing with the formal solution of the exact phase coexistence problem. Assume that the parent $`\rho ^{(0)}(\sigma )`$ has separated into $`p`$ phases numbered by $`\alpha =1\mathrm{}p`$. The condition (51), which follows from equality of the chemical potentials $`\mu (\sigma )`$ in all phases, implies that we can write their density distributions $`\rho ^{(\alpha )}(\sigma )`$ as
$$\rho ^{(\alpha )}(\sigma )=\stackrel{~}{R}(\sigma )\mathrm{exp}\left(\underset{i}{}\lambda _i^{(\alpha )}w_i(\sigma )\right)$$
(61)
for some function $`\stackrel{~}{R}(\sigma )`$. If phase $`\alpha `$ occupies a fraction $`v^{(\alpha )}`$ of the system volume, particle conservation $`_\alpha v^{(\alpha )}\rho ^{(\alpha )}(\sigma )=\rho ^{(0)}(\sigma )`$ then gives
$$\stackrel{~}{R}(\sigma )=\frac{\rho ^{(0)}(\sigma )}{_\alpha v^{(\alpha )}\mathrm{exp}\left(_i\lambda _i^{(\alpha )}w_i(\sigma )\right)},\rho ^{(\alpha )}(\sigma )=\rho ^{(0)}(\sigma )\frac{\mathrm{exp}\left(_i\lambda _i^{(\alpha )}w_i(\sigma )\right)}{_\beta v^{(\beta )}\mathrm{exp}\left(_i\lambda _i^{(\beta )}w_i(\sigma )\right)}$$
(62)
If there are $`K`$ moment densities $`\rho _i`$, $`i=1\mathrm{}K`$, then this exact solution is parameterized by $`(p1)(K+1)`$ independent parameters: $`(p1)K`$ parameters $`\lambda _i^{(\alpha )}`$ (noting that the $`\lambda _i`$ of one phase can be fixed arbitrarily), and $`p1`$ parameters $`v^{(\alpha )}`$ (noting that one phase volume is fixed by the constraint $`_\alpha v^{(\alpha )}=1`$). Comparing (61) with (51), one sees that the $`K`$ quantities
$`T\lambda _i^{(\alpha )}+\stackrel{~}{\mu }_i^{(\alpha )}(i=1\mathrm{}K)`$
must be the same in all phases $`\alpha `$; the same is true for the pressures $`\mathrm{\Pi }^{(\alpha )}`$, and this gives the required total number of $`(p1)(K+1)`$ constraints. This formally defines the exact solution of the phase coexistence problem for truncatable systems. Its practical value is limited by the difficulties of finding the solution numerically, as pointed out in Sec. I; see also Sec. IV. There is also no interpretation of the result in terms of a free energy depending on a small number of densities.
Nevertheless, eq. (62) is useful for a comparison with the solution provided by the moment free energy method. From (61), the exact coexisting phases are all members of a family of density distributions of the general form (7); this is simply a consequence of the requirement of equal chemical potentials (51). This โexact coexistence familyโ has an โeffectiveโ prior $`\stackrel{~}{R}(\sigma )\rho ^{(0)}(\sigma )`$ and is therefore different from the โoriginalโ family (42). The moment free energy only gives us access to distributions from the original family, not the exact coexistence family, and therefore cannot yield exact solutions for phase coexistence (beyond its onset). Fig. 2 illustrates this point explicitly in the simplified context of a bidisperse system.
It is now easy to see, however, that โ as stated in Sec. II A โ the exact solution can be approached to arbitrary precision by including extra moment densities in the moment free energy. (This leaves the exactness of spinodals, critical points, cloud-points and shadows unaffected, because none of our arguments excluded a null dependence of $`\stackrel{~}{f}`$ on certain of the $`\rho _i`$.) Indeed, by adding further moment densities one can indefinitely extend the family (42) of density distributions, thereby approaching with increasing precision the actual distributions in all phases present; this yields phase diagrams of ever-refined accuracy.
The roles of the โoriginalโ moment densities (those appearing in the excess free energy) and the extra ones are quite different, however. To see this, note from (45) that equality among phases of the moment chemical potentials implies that, for the extra moments only, the corresponding Lagrange multipliers must themselves be equal in all coexisting phases. (This is because there is, by construction, no excess part to the โextraโ chemical potentials.) We can therefore drop the phase index on these Lagrange multipliers and write the density distribution in phase $`\alpha `$ as
$$\rho ^{(\alpha )}(\sigma )=\rho ^{(0)}(\sigma )\mathrm{exp}\left(\underset{\mathrm{extra}i}{}\lambda _iw_i(\sigma )\right)\mathrm{exp}\left(\underset{\mathrm{original}i}{}\lambda _i^{(\alpha )}w_i(\sigma )\right)$$
(63)
Comparing with (7) and (61), we see that the extra Lagrange multipliers can be thought of as providing a โflexible priorโ $`R(\sigma )`$, allowing a better approximation to the effective prior $`\stackrel{~}{R}(\sigma )`$ required by particle conservation. The โtuningโ of the prior with these extra Lagrange multipliers also has an important effect on the number of coexisting phases that can be found by the moment method (see Sec. III G).
### F Global and local stability
Most numerical algorithms for phase coexistence calculations, including ours, initially proceed by finding a solution to the equilibrium conditions of equal chemical potentials and pressures in all phases, rather than by a direct minimization of the total free energy of the system. It is then crucial to verify whether this solution is stable, both locally (i.e., with respect to small fluctuations in the compositions of the phases) and globally with respect to splitting into a larger number (or different) phases (see Fig. 3). Global stability is a particularly important issue in our context because there is in principle no limit on the number of coexisting phases in a polydisperse system.
A useful tool for stability calculations is the โtangent plane distanceโ . Let us first define this generically for a system with (a vector of) densities $`๐`$, free energy density $`f(๐)`$ and chemical potentials $`๐(๐)=f(๐)`$; this notation is the same as in Sec. III A. Assume we have found a candidate โphase splitโ, that is a collection of $`p`$ phases $`๐^{(\alpha )}`$ ($`\alpha =1\mathrm{}p`$) which satisfy the phase equilibrium conditions of equal chemical potentials $`๐=๐^{(\alpha )}`$ and pressures $`\mathrm{\Pi }=\mathrm{\Pi }^{(\alpha )}`$. They occupy fractions $`v^{(\alpha )}`$ of the total volume, and the overall density distribution is thus $`๐_{\mathrm{tot}}=_\alpha v^{(\alpha )}๐^{(\alpha )}`$. Note that we do not yet assume that the lever rule is satisfied, i.e., that $`๐_{\mathrm{tot}}=๐^{(0)}`$; as explained above, this equality will generally not hold for phase splits calculated from the moment free energy. We now define โglobal tangent plane (TP) stabilityโ for such a phase split as the property that there is no other phase split (i.e., no other tangent plane) that gives a lower total free energy for the same overall density distribution $`๐_{\mathrm{tot}}`$. In more intuitive language, this means that if we were to put the phases $`๐^{(\alpha )}`$ into contact with each other, the resulting system would be thermodynamically stable; neither the composition nor the number of phases would change over time. Note, however, that since in general $`๐_{\mathrm{tot}}=_\alpha v^{(\alpha )}๐^{(\alpha )}`$ is not equal to $`๐^{(0)}`$, the phase split that is globally TP-stable need not accurately reflect the number and composition of phases into which the parent $`๐^{(0)}`$ would actually split under the chosen thermodynamic conditions.
To define the tangent plane distance (TPD), note first that by virtue of the coexistence conditions, all phases $`๐^{(\alpha )}`$ lie on a tangent plane to the free energy surface. Points $`(๐,f)`$ on this tangent plane obey the equation $`f๐๐+\mathrm{\Pi }=0`$, with $`๐`$ and $`\mathrm{\Pi }`$ the chemical potentials and pressure common to all phases. For a generic phase with density distribution $`๐`$ and free energy $`f(๐)`$, the same expression will have a nonzero value which measures how much โbelowโ or โaboveโ the tangent plane it lies. This defines the TPD
$$t(๐)=f(๐)๐^{(\alpha )}๐+\mathrm{\Pi }^{(\alpha )}.$$
(64)
Here we have added the superscript $`\alpha `$ to emphasize that the chemical potential and osmotic pressure used in the calculation of the TPD are those of the calculated phase equilibrium (and hence of any of the participating phases $`\alpha `$), rather than those of the test phase $`๐`$. It is then clear intuitively โ and can be shown more formally โ that the calculated phase coexistence is globally TP-stable if the TPD is non-negative everywhere. Geometrically, this simply means that no part of the free energy surface must protrude beneath the tangent plane; otherwise the total free energy of the system could be lowered by constructing a new tangent plane that touches the protruding piece. Global TP-stability of course encompasses local stability; the latter simply corresponds to the requirement that the TPD be a local minimum (with the value $`t=0`$) at each of the phases in the candidate solution.
When verifying global TP-stability, it is obviously sufficient to check the value of the TPD at all of its stationary points. By differentiating (64) w.r.t. $`๐`$, one sees that at these points, the chemical potentials $`๐(๐)`$ are the same as in the calculated coexisting phases. For our truncatable polydisperse systems, it then follows from eq. (51) that we only need to consider the TPD for test phases $`\rho (\sigma )`$ which are in the same family (7) as the calculated coexisting phases. This is a crucial point: Even though global TP-stability is a statement about stability in the infinite-dimensional space of density distributions $`\rho (\sigma )`$, it can be checked by only considering density distributions from a $`K`$-dimensional family. In the (generally hypothetical) case of an exactly calculated phase split, this family would be (62), with the effective prior $`\stackrel{~}{R}(\sigma )`$ . For a phase split calculated from the moment free energy, it is the family (42) with the parent as prior, with any Lagrange multipliers for extra moments fixed to their values in the calculated coexisting phases. The TPD of such a test phase is (using (41) for the exact free energy and (48) for the chemical potentials $`\mu ^{(\alpha )}(\sigma )`$ of the coexisting phases)
$`t[\rho (\sigma )]`$ $`=`$ $`f[\rho (\sigma )]{\displaystyle ๐\sigma \mu ^{(\alpha )}(\sigma )\rho (\sigma )}+\mathrm{\Pi }^{(\alpha )}`$ (65)
$`=`$ $`\stackrel{~}{f}(\rho _i)+{\displaystyle ๐\sigma \rho (\sigma )\left\{T\left[\mathrm{ln}\rho (\sigma )1\right]T\mathrm{ln}\rho ^{(\alpha )}(\sigma )\underset{i}{}\stackrel{~}{\mu }_i^{(\alpha )}w_i(\sigma )\right\}}+\mathrm{\Pi }^{(\alpha )}`$ (66)
$`=`$ $`\stackrel{~}{f}(\rho _i)+T{\displaystyle ๐\sigma \rho (\sigma )\left[\mathrm{ln}\frac{\rho (\sigma )}{\rho ^{(0)}(\sigma )}1\right]}{\displaystyle \underset{i}{}}\left[\lambda _i^{(\alpha )}+\stackrel{~}{\mu }_i^{(\alpha )}\right]\rho _i+\mathrm{\Pi }^{(\alpha )}`$ (67)
$`=`$ $`f_\mathrm{m}(\rho _i){\displaystyle \underset{i}{}}\mu _i^{(\alpha )}\rho _i+\mathrm{\Pi }^{(\alpha )}.`$ (68)
A comparison with (64) shows that this is identical to the TPD that one would derive from the moment free energy alone. Here again, the underlying polydisperse nature of the problem can therefore be disregarded once the moment free energy has been obtained. In summary, one can determine whether a phase split calculated from the moment free energy is globally TP-stable (which means that phases of the predicted volumes and compositions, would, if placed in contact, indeed coexist) using only the TPD derived from the moment free energy; the same is trivially true of the weaker requirement of local stability .
Recall, however, that the overall density distribution $`\rho _{\mathrm{tot}}(\sigma )`$ for a phase split found from the moment free energy only has the same moment densities $`\rho _i^{(0)}`$ as the parent $`\rho ^{(0)}(\sigma )`$, but differs in other details (the transverse degrees of freedom). Global TP-stability thus guarantees that such an approximate phase split is thermodynamically stable, but does not imply that it is identical (in either number or composition of phases) to the exact one. Nevertheless, progressively increasing the number of extra moment densities in the moment free energy will make $`\rho _{\mathrm{tot}}(\sigma )`$ a progressively better approximation to $`\rho ^{(0)}(\sigma )`$, and so the exact phase split, with the correct number of phases, must eventually be recovered to arbitrary accuracy; see Fig. 4 for an illustration.
### G Geometry in density distribution space
In the above discussion of the properties of the moment free energy, we have focused on obtaining the phase behavior of a system with a given parent distribution $`\rho ^{(0)}(\sigma )`$. This is the point of view most relevant for practical applications of the method, and the rest of this Section is not essential for understanding such applications. Nevertheless, from a theoretical angle, it is also interesting to consider the global geometry of the space of all density distributions $`\rho (\sigma )`$, without reference to a specific parent: the exact free energy (41) of the underlying (truncatable) model induces two-phase tie lines and multi-phase coexistence regions in this space, and one is led to ask how the moment free energy encodes these properties.
As pointed out in Sec. II A, the definition of the moment free energy depends on a prior $`R(\sigma )`$ and represents the properties of systems with density distributions $`\rho (\sigma )`$ in the corresponding maximum entropy family (7). Instead of identifying $`R(\sigma )=\rho ^{(0)}(\sigma )`$, we now allow a general prior $`R(\sigma )`$. Conceptually, it then makes sense to associate the moment free energy with the family (7) rather than the specific prior. This is because any density distribution from (7) can be chosen as prior, without changing the moment free energy (apart from irrelevant linear terms in the moments $`\rho _i`$), or the identity of the remaining family members. The construction of the moment free energy thus partitions the space of all $`\rho (\sigma )`$ into (an infinite number of) families (7); different families give different moment free energies that describe the thermodynamics โwithin the familyโ. This procedure gives meaningful results because, as shown at the end of Sec. III A, coexisting phases are always members of the same family. By considering the entire ensemble of families (7) and their corresponding moment free energies, one can thus in principle recover the exact geometry of the density distribution space and the phase coexistences within it. Note that this includes regions with more than $`K+1`$ phases, even when the excess (and thus moment) free energy depends on only $`K`$ moment densities; see Fig. 4. Consistent with Gibbsโ phase rule, however, the families for which this occurs are exceptional: they occupy submanifolds (of measure zero) in the space of all families. Accordingly, to find the families on which such โsuper-Gibbsโ multi-phase coexistences occur, the corresponding prior must be very carefully tuned. Indeed, the probability of finding such priors โaccidentallyโ, without either adding extra moments to the moment free energy description or solving the exact phase coexistence problem, is zero. Generically, one needs to retain at least $`n`$ moment densities in the moment free energy to find $`n+1`$ phases in coexistence. From (63), this corresponds to having $`nK`$ parameters available for tuning the location of the family (or equivalently, its prior) in density distribution space. But note that, once the extra moments have been introduced, the required tuning need not be done โby handโ: it is achieved implicitly by requiring the lever rule to be obeyed not only for the $`K`$ original moment densities, but also for the $`nK`$ extra ones. (The solution of the phase equilibrium conditions thus effectively proceeds in an $`n`$-dimensional space; but once a solution has been found, its (global TP-)stability can still be checked by computing the TPD only within a $`K`$-dimensional family of distributions. See Sec. III F.)
As pointed out above, phase splits calculated from the moment free energy allow violations of the lever rule. In the global view, this fact also has a simple geometric interpretation: the families (7) are (generically) curved. In other words, for two distributions $`\rho ^{(1)}(\sigma )`$ and $`\rho ^{(2)}(\sigma )`$ from the same family, the straight line $`\rho (\sigma )=ฯต\rho ^{(1)}(\sigma )+(1ฯต)\rho ^{(2)}(\sigma )`$ connecting them lies outside the family. More generally, if $`p`$ coexisting phases $`\rho ^{(\alpha )}(\sigma )`$ have been identified from the moment free energy, then the overall density distribution $`\rho _{\mathrm{tot}}(\sigma )=_\alpha v^{(\alpha )}\rho ^{(\alpha )}(\sigma )`$ is different from the parent $`\rho ^{(0)}(\sigma )`$; geometrically, it lies on the hyperplane that passes through the phases $`\rho ^{(\alpha )}(\sigma )`$.
## IV Practical implementation of the moment method
The application of the moment free energy method to the calculation of spinodal and critical points is straighforward using conditions (58), and (60) or (D3), respectively, and is further illustrated in Sec. V below. We therefore focus in this Section on phase coexistence calculations.
Recall that in the moment approach, each phase $`\alpha `$ is parameterized by Lagrange multipliers $`\lambda _i^{(\alpha )}`$ for the original moments (the ones appearing in the excess free energy of the system) and the fraction $`v^{(\alpha )}`$ of system volume that it occupies. If extra moments are used, there is one additional Lagrange multiplier $`\lambda _i`$ for each of them; these are common to all phases. These parameters have to be chosen such that the pressure (47) and the moment chemical potentials $`\mu _i`$ given by (45) are equal in all phases. Furthermore, the (fractional) phase volumes $`v^{(\alpha )}`$ have to sum to one, and the lever rule has to be satisfied for all moments (both original and extra):
$`{\displaystyle \underset{\alpha }{}}v^{(\alpha )}\rho _i^{(\alpha )}=\rho _i^{(0)}.`$
In a system with $`K`$ original moments that is being studied using an $`n`$-moment free energy (i.e., with $`nK`$ extra moments), and for $`p`$ coexisting phases, one has $`p(K+1)+nK=(p1)(K+1)+n+1`$ parameters and as many equations. Starting from a suitable initial guess, these can, in principle, be solved by a standard algorithm such as Newton-Raphson . Generating an initial point from which such an algorithm will converge, however, is a nontrivial problem, especially when more than two phases coexist.
To simplify this task, we work with a continuous control parameter such as temperature, density $`\rho _0`$ of the parent phase, or interaction parameter $`\chi `$ for polymers. Taking the latter case as an example, we start the calculation at a small value of $`\chi `$ where we are sure to be in a single-phase region; thus the parent phase under consideration is stable. The basic strategy is then to increment $`\chi `$ and detect potential new phases as we go along. At each step, the phase equilibrium conditions are solved by Newton-Raphson for the current number of phases. Then we check for local stability of the solution by calculating the Hessian of the TPD around each of the coexisting phases and verifying that it is positive definite. If an instability (i.e., a negative eigenvalue of the Hessian) is found, we search for local minima starting from points displaced either way along the instability direction. If two new local minima are found in this way, we add them to list of phases and delete the unstable phase; if only one new local minimum is uncovered, we add this but retain the old phase. Finally, we check for global (TP-)stability. As explained in Sec. III F, this involves scanning all โtestโ phases from the same maximum entropy family for possible negative values of the TPD. However, the test phase will have the same extra Lagrange multipliers, and is therefore parameterized in terms of the $`\lambda _i`$ for the original moments only. The extra Lagrange multipliers are held fixed, so their number is irrelevant for the computational cost of the stability check. Put differently, within our algorithm the precision of the coexistence curves should depend on the total number of moments retained, $`n`$, rather than the number of moments in the excess free energy, $`K`$. Computational effort, on the other hand, is dominated by the global stability check, and thus mainly sensitive to $`K`$ (the dimension of the space to be searched for new phases), rather than $`n`$.
This is a substantial efficiency gain, but despite it, an exhaustive search for local TPD minima over the $`K`$-dimensional space of original Lagrange multipliers is unrealistic except for $`K=1`$. We use instead a Monte-Carlo-type algorithm to sample the TPD at representative points. If any negative values of the TPD are encountered, we choose the smallest such value, find the nearest local minimum of the TPD and add this point to the list of phases. If any new phases have been found during the (local and global) stability checks, we assign each of them a default phase volume ($`v=0.1`$, say), reduce the phase volumes of the old phases accordingly, update the number of phases and loop back to the Newton-Raphson solution of the phase equilibrium conditions, using the current list of phases as an initial guess. This process is repeated until the calculated solution is found to be stable. Our implementation of this basic scheme also contains some additional elements (such as an adaptive choice of the stepsize for the control parameter $`\chi `$, and checks for very small phase volumes) which are useful near points where new phases appear or old ones vanish.
Comparing our own approach, which is based on the moment free energy with extra moment densities as just outlined, to (62), which is the exact solution for truncatable systems, we see that the former actually uses more parameters (essentially one phase-independent Lagrange multiplier per extra moment) to represent the solution. At first sight, this may appear counter-productive. However, the Lagrange multipliers $`\lambda _i^{(\alpha )}`$ and the (fractional) phase volumes $`v^{(\alpha )}`$ are much less strongly coupled in the moment free energy approach. The phase volumes are only determined by the lever rule, while in the exact solution they โfeed backโ into the effective prior $`\stackrel{~}{R}(\sigma )`$ and therefore into the equilibrium conditions of equal chemical potentials and pressures. In the examples studied below, we have found that the numerical advantages of this decoupling (in terms of stability, robustness and convergence of our algorithm) can easily outweigh the larger number of parameters that it requires.
As with any numerical algorithm, it is important to develop robust criteria by which the convergence of the solution can be judged. This has a novel aspect, in the moment method, since violations of the lever rule are allowed: alongside normal numerical convergence criteria one needs a method for deciding whether the effect of these on the predicted phase behaviour is significant. We develop appropriate criteria in Sec. V B in the context of a specific example. The basic ideas is that since any state of phase coexistence predicted by the method represents the exact behaviour of some parent, then so long as this parent is close enough to the true one, the predicted behaviour will lie within the range of uncertainty that arises anyway, from not knowing the true parent to arbitrary experimental precision.
## V Examples
We now illustrate how the moment method is applied and demonstrate its usefulness for several examples. The first two (Flory-Huggins theory for length-polydisperse homopolymers and dense chemically polydisperse copolymers, respectively) contain only a single moment density in the excess free energy and are therefore particularly simple to analyse and visualize. In the third example (chemically polydisperse copolymers in a polymeric solvent), the excess free energy depends on two moment densities and this will gives us the opportunity to discuss the appearance of more complex phenomena such as tricritical points.
### A Homopolymers with length polydispersity
Let us start with the simple but well studied example of polydisperse Flory-Huggins theory . One considers a system of homopolymers with a distribution of chain lengths; the polydisperse feature $`\sigma `$ is simply the chain length $`L`$, i.e., the number of monomers in each chain. (We treat this as a continuous variable.) The density distribution $`\rho (L)`$ then gives the number density of chains as a function of $`L`$. We choose the segment volume $`a^3`$ as our unit of volume, making $`\rho (L)`$ dimensionless. The volume fraction occupied by the polymer is then simply the first moment of this distribution, $`\varphi \rho _1=๐LL\rho (L)`$. Within Flory-Huggins theory, the free energy density is (in units such that $`k_\text{B}T=1`$)
$$f=๐L\rho (L)[\mathrm{ln}\rho (L)1]+(1\rho _1)\mathrm{ln}(1\rho _1)+\chi \rho _1(1\rho _1)$$
(69)
where the Flory $`\chi `$-parameter plays essentially the role of an inverse temperature. Before analysing this further, we note that for a bidisperse system with chain lengths $`L_1`$ and $`L_2`$ and number densities $`r_1`$ and $`r_2`$, the corresponding expression would be
$$f=r_1(\mathrm{ln}r_11)+r_2(\mathrm{ln}r_21)+(1\rho _1)\mathrm{ln}(1\rho _1)+\chi \rho _1(1\rho _1),\rho _1=L_1r_1+L_2r_2$$
(70)
This free energy, with $`L_1=10`$ and $`L_2=20`$, was used to generate the examples shown in Figs. 1 and 2.
Returning now to (69), we note first that the excess free energy
$$\stackrel{~}{f}=(1\rho _1)\mathrm{ln}(1\rho _1)+\chi \rho _1(1\rho _1)$$
(71)
depends only on the moment density $`\rho _1`$. If no extra moments are used, the moment free energy is therefore a function of a single density variable, $`\rho _1`$. Let us work out its construction explicitly for the case of a parent phase with a Schulz-distribution of lengths, given by
$$\rho ^{(0)}(L)=\rho _0^{(0)}\frac{1}{\mathrm{\Gamma }(a)b^a}L^{a1}e^{L/b}$$
(72)
Here $`a`$ is a parameter that determines how broad or peaked the distribution is; it is conventionally denoted by $`\alpha `$, but we choose a different notation here to prevent confusion with the phase index used in the general discussion so far. The chain number density of the parent is $`\rho _0^{(0)}`$, and the normalized first moment $`m_1^{(0)}=\rho _1^{(0)}/\rho _0^{(0)}`$ is simply the (number) average chain length
$$L_Nm_1^{(0)}\frac{\rho _1^{(0)}}{\rho _0^{(0)}}=ab.$$
(73)
The parent is thus parameterized in terms of $`a`$, $`\rho _0^{(0)}`$ and $`\rho _1^{(0)}`$ (or $`b`$). The density distributions in the family (42) are given by
$$\rho (L)=\rho ^{(0)}(L)e^{\lambda _1L}=\rho _0^{(0)}\frac{1}{\mathrm{\Gamma }(a)b^a}L^{a1}e^{(b^1\lambda _1)L}$$
(74)
and the zeroth and first moment densities are easily worked out, for members of this family, to be
$`\rho _0=\rho _0^{(0)}\left({\displaystyle \frac{1}{1b\lambda _1}}\right)^a,\rho _1=\rho _1^{(0)}\left({\displaystyle \frac{1}{1b\lambda _1}}\right)^{a+1}.`$
Because the family only has a single parameter $`\lambda _1`$, these two moment densities are of course related to one another:
$$\rho _0=c\rho _1^{a/(a+1)},c=\rho _0^{(0)}\left(\rho _1^{(0)}\right)^{a/(a+1)}.$$
(75)
Now we turn to the moment entropy, given by (40): $`s_\mathrm{m}=\rho _0\lambda _1\rho _1`$, and to be considered as a function of $`\rho _1`$. After a little algebra (and eliminating $`b`$ in favor of $`\rho _1^{(0)}`$) one finds
$$s_\mathrm{m}=\rho _0^{(0)}\left[(1+a)\left(\frac{\rho _1}{\rho _1^{(0)}}\right)^{a/(a+1)}a\frac{\rho _1}{\rho _1^{(0)}}\right]$$
(76)
The last term is linear in $`\rho _1`$ and can be disregarded for the calculation of phase equilibria. This then gives the following simple result for the moment free energy
$$f_\mathrm{m}=(a+1)c\rho _1^{a/(a+1)}+\stackrel{~}{f}.$$
(77)
(In conventional polymer notation, this result would read $`f_\mathrm{m}=(\alpha +1)c\varphi ^{\alpha /(\alpha +1)}+\stackrel{~}{f}`$.) From this we can now obtain the spinodal condition, for example, which identifies the value of $`\chi `$ where the parent becomes unstable. The general criterion (53) simplifies in our case of a single moment density to
$$\frac{d^2f_\mathrm{m}}{d\rho _1^2}|_{\rho _1=\rho _1^{(0)}}=0\frac{1}{1\rho _1^{(0)}}2\chi +\frac{a}{a+1}\frac{\rho _0^{(0)}}{(\rho _1^{(0)})^2}=0$$
(78)
The same condition for a phase with general density distribution $`\rho (L)`$ โ rather than our specific Schulz parent $`\rho ^{(0)}(L)`$ โ follows from (58) as
$$\frac{1}{1\rho _1}2\chi +\frac{1}{\rho _2}$$
(79)
As expected, this becomes equivalent to (78) for parents of the Schulz form (72), which obey $`\rho _2^{(0)}=\rho _0^{(0)}a(a+1)b^2`$ and eq. (73). Note that in standard polymer notation, eq. (79) would be written as
$$\frac{1}{1\varphi }2\chi +\frac{1}{L_W\varphi }$$
(80)
where $`L_W=\rho _2/\rho _1`$ is the weight average chain length.
The critical point condition is obtained similarly. For a single density variable, eq. (59) reduces to
$$\frac{d^3f_\mathrm{m}}{d\rho _1^3}|_{\rho _1=\rho _1^{(0)}}=0\frac{1}{(1\rho _1^{(0)})^2}\frac{a(a+2)}{(a+1)^2}\frac{\rho _0^{(0)}}{(\rho _1^{(0)})^3}=0$$
(81)
For a general distribution, the critical point criterion (60) becomes instead
$`{\displaystyle \frac{\rho _3}{\rho _2^3}}+{\displaystyle \frac{d^3\stackrel{~}{f}}{d\rho _1^3}}=0`$
and inserting the Flory-Huggins excess free energy (71) gives
$$\frac{1}{(1\rho _1)^2}\frac{\rho _3}{\rho _2^3}=0$$
(82)
Again, this simplifies to (81) for Schulz parents, for which $`\rho _3^{(0)}=\rho _0^{(0)}a(a+1)(a+2)b^3`$. The traditional statement of (82), in terms of $`\varphi \rho _1`$, $`L_W`$ and the $`Z`$-average chain length $`L_Z=\rho _3/\rho _2`$ reads
$$\frac{1}{(1\varphi )^2}\frac{L_Z}{L_W^2}\frac{1}{\varphi ^2}=0.$$
(83)
Fig. 5 shows the moment free energy (77) for a given parent and different values of $`\chi `$. The value of $`\chi `$ for which there is a double tangent at the point $`\rho _1=\rho _1^{(0)}`$ representing the parent gives the cloud point; the second tangency point gives the polymer volume fraction $`\rho _1`$ in the coexisting phase, hence the shadow point. Repeating this procedure for different values of $`\rho _1^{(0)}`$ (while maintaining $`L_N=\rho _1^{(0)}/\rho _0^{(0)}`$ and $`a`$ constant), one obtains the full CPC and shadow curve for a given normalized parent length distribution $`n(L)=\rho ^{(0)}(L)/\rho _0^{(0)}`$.
We now consider the properties of the moment free energy with the (chain number) density $`\rho _0`$ retained as an extra moment. This provides additional geometrical insight into the properties of polydisperse chains (while for the numerical determination of the CPC and shadow curve, the above one-moment free energy is preferable). To construct the two-moment free energy, we proceed as before. The family (42) is now
$$\rho (L)=\rho ^{(0)}(L)e^{\lambda _0+\lambda _1L}=\rho _0^{(0)}\frac{1}{\mathrm{\Gamma }(a)b^a}L^{a1}e^{\lambda _0(b^1\lambda _1)L}$$
(84)
with zeroth and first moment densities
$`\rho _0=\rho _0^{(0)}e^{\lambda _0}\left({\displaystyle \frac{1}{1b\lambda _1}}\right)^a,\rho _1=\rho _1^{(0)}e^{\lambda _0}\left({\displaystyle \frac{1}{1b\lambda _1}}\right)^{a+1}.`$
The relations can be inverted to express the Lagrange multipliers $`\lambda _0`$ and $`\lambda _1`$ in terms of the moment densities as
$`\lambda _0`$ $`=`$ $`(a+1)\mathrm{ln}{\displaystyle \frac{\rho _0}{\rho _0^{(0)}}}a\mathrm{ln}{\displaystyle \frac{\rho _1}{\rho _1^{(0)}}}`$
$`\lambda _1`$ $`=`$ $`{\displaystyle \frac{1}{b}}\left(1{\displaystyle \frac{\rho _0/\rho _0^{(0)}}{\rho _1/\rho _1^{(0)}}}\right)`$
and one obtains the moment entropy
$$s_\mathrm{m}=\rho _0\lambda _0\rho _0\lambda _1\rho _1=(a+1)\rho _0\mathrm{ln}\frac{\rho _0}{\rho _0^{(0)}}+a\rho _0\mathrm{ln}\frac{\rho _1}{\rho _1^{(0)}}+(a+1)\rho _0a\frac{\rho _0^{(0)}}{\rho _1^{(0)}}\rho _1.$$
(85)
As expected, this reduces to (76) for systems with density distributions from our earlier one-parameter family (74), where $`\rho _0`$ can be expressed as a function of $`\rho _1`$ according to (75). With both $`\rho _0`$ and $`\rho _1`$ retained as moment densities in the moment free energy, a number of linear terms in (85) can be dropped, giving the final result
$$f_\mathrm{m}=(a+1)\rho _0\mathrm{ln}\rho _0a\rho _0\mathrm{ln}\rho _1+\stackrel{~}{f}.$$
(86)
Note that the dependence on the parent distribution is now only through $`a`$. This can be understood from the general discussion in Sec. III G: The family (84) of density distributions now contains all Schulz-distributions with the given $`a`$, and the moment free energy is insensitive to which member of this family (specified by $`\rho _0^{(0)}`$ and $`\rho _1^{(0)}`$) is used as the parent.
The result (86) can of course also be obtained via the combinatorial method, as follows. The normalized Schulz parent distribution is given by $`n^{(0)}(L)=[\mathrm{\Gamma }(a)b^a]^1L^{a1}e^{bL}`$. The cumulant generating function is thus $`h(\lambda _1)=a\mathrm{ln}(1b\lambda _1)`$, giving $`m_1=h/\lambda _1=ab/(1b\lambda _1)`$. Dropping constants and linear terms in the average chain length $`m_1`$ (which do not affect the phase behavior), the generalized entropy of mixing per particle, $`s_{\mathrm{mix}}=h\lambda _1m_1`$ then becomes $`s_{\mathrm{mix}}=a\mathrm{ln}m_1`$. Adding the ideal gas term and the excess part of the free energy, we thus find the moment free energy density
$$f_\mathrm{m}=\rho _0\mathrm{ln}\rho _0a\rho _0\mathrm{ln}(\rho _1/\rho _0)+\stackrel{~}{f}$$
(87)
in agreement with (86); it is a simple two component free energy. In the conventional notation, eq. (87) would read
$`f_\mathrm{m}=\rho \mathrm{ln}\rho \alpha \rho \mathrm{ln}(\varphi /\rho )+(1\varphi )\mathrm{ln}(1\varphi )+\chi \varphi (1\varphi ).`$
The spinodal curve (SC) and critical point (CP) condition may now be calculated from eq. (87), using either the methods outlined in Sec. III A or the more traditional determinant conditions (see also App. D). One finds
$`{\displaystyle \frac{1}{1\rho _1}}2\chi +{\displaystyle \frac{a}{a+1}}{\displaystyle \frac{\rho _0}{\rho _1^2}}=0,`$ $`(\mathrm{SC})`$ (88)
$`{\displaystyle \frac{1}{(1\rho _1)^2}}{\displaystyle \frac{a(a+2)}{(a+1)^2}}{\displaystyle \frac{\rho _0}{\rho _1^3}}=0.`$ $`(\mathrm{CP})`$ (89)
Evaluated at the parent, these conditions are identical to eqs. (78) and (81) โ which were derived from the one-moment free energy โ as they must be. But, as explained at the end of Sec. III A, they also hold more generally for all systems with Schulz density distributions with the given value of $`a`$; see the remarks after eq. (86). Of particular interest among these systems are those that differ from the parent only in their overall density while having the same normalized length distribution $`n(L)=n^{(0)}(L)`$. They form the โdilution lineโ; their number averaged chain length satisfies $`m_1=m_1^{(0)}=๐LLn^{(0)}(L)`$, which translates to $`\rho _1=L_N\rho _0`$ (where $`L_Nm_1^{(0)}=ab`$ is, as before, the number average chain length of the parent). Inserting this โdilution line constraintโ into the above equations, we obtain as the spinodal and critical point conditions for systems on the dilution line
$`{\displaystyle \frac{1}{1\rho _1}}2\chi +{\displaystyle \frac{a}{a+1}}{\displaystyle \frac{1}{L_N\rho _1}}=0,`$ $`(\mathrm{SC})`$ (90)
$`{\displaystyle \frac{1}{(1\rho _1)^2}}{\displaystyle \frac{a(a+2)}{(a+1)^2}}{\displaystyle \frac{1}{L_N\rho _1^2}}=0.`$ $`(\mathrm{CP})`$ (91)
Again, agreement with the general results (80) and (83) is easily verified by noting that for Schulz length distributions, $`L_W=\rho _2/\rho _1=(a+1)b=L_N(a+1)/a`$ and $`L_Z=\rho _3/\rho _2=(a+2)b=L_N(a+2)/a`$ for the weight and $`Z`$-average chain lengths, respectively.
The above discussion focussed on spinodals, critical points, and cloud and shadow curves, all of which are found exactly by the moment approach. (As usual, โexactโ is subject to the assumed validity of the original, truncatable free energy expression (71).) Within the moment approach, we have also calculated the full phase behavior in the $`(\rho _1,\rho _0)`$ plane. This was done numerically from the two-moment free energy (87), facilitated by a partial Legendre transformation from $`\rho _0`$ to the associated chemical potential $`\mu _0`$, which can be performed analytically in this case (and effectively brings one back to the one-moment free energy (77)). Fig. 6(a)โ(c) shows binodal curves, tielines, spinodal curves and critical points in the $`(\rho _1,\rho _0)`$ plane, for three values of $`\chi `$. These plots give an appealing geometric insight into the problem of phase separation in this system. For instance, the slope of the tielines indicates a size partitioning effect: the more dense phases are enriched in long chains.
The heavy line in Fig. 6(a)โ(c) is the dilution line constraint $`m_1=m_1^{(0)}`$ or $`\rho _1=L_N\rho _0`$. As discussed already, we are interested mainly in systems whose mean composition lies on this line. (With $`\chi `$ as an extra variable, this line becomes a plane ($`\rho _1=L_N\rho _0,\chi `$) in the space $`(\rho _1,\rho _0,\chi )`$.) Not all of the phase behavior shown in Fig. 6 is then accessible. The extremities of phase separation on the $`\rho _1=L_N\rho _0`$ line are points where phase separation just starts to occur, and the locus of these points in the $`(\rho _1=L_N\rho _0,\chi )`$ plane defines the cloud curve, which bounds the phase separation region. This is shown in Fig. 6(d). This plot also includes the spinodal stability curve and the critical point from eqs. (90) and (91). The spinodal curve and the cloud curve touch at the critical point, which no longer lies at the minimum of either. This distorted behavior is a well known feature of polydisperse systems. Here it is seen to be due to the way that *regular* phase behavior in $`(\rho _1,\rho _0,\chi )`$ space is cut through by the dilution line constraint. Note that only two moments are needed to understand this qualitative effect, and that Fig. 6 gives direct geometrical insight into it, which would be very hard to extract from any exact solution based on (61) and (62).
The compositions of the phases that just start to appear as the phase separation region is entered do not in general lie on the dilution line $`\rho _1=L_N\rho _0`$. These phases are the โshadowsโ; they lie at the other ends of the cloud point tielines in Fig. 6(a)โ(c). This is illustrated for $`\chi =0.62`$ in Fig. 7. The shadow phase compositions may be projected onto the ($`\rho _1=L_N\rho _0`$, $`\chi `$) plane to give shadow curves, but the projection is not unique. The shadow curve obtained by ignoring the value of $`\rho _0`$ and projecting onto the $`\rho _1`$-axis, for example, is shown in Fig. 8(a). That obtained by ignoring the value of $`\rho _1`$ and projecting onto the $`\rho _0`$-axis is shown in Fig. 8(b). The different shapes are due to the fact that the locus of shadow phase compositions in general only intersects the $`(\rho _1=L_N\rho _0,\chi )`$ plane at a single pointโthe critical point.
In conclusion, we emphasize again that all the curves shown in Fig. 6(d) and Fig. 8 are exact, even though they have been constructed from the phase behavior shown in Fig. 6(a)โ(c) which is itself a projection of the true phase behavior of the fully polydisperse system.
### B Copolymer with chemical polydispersity
Next, we apply the moment free energy formalism to random AB-copolymers in the melt or in solution . For simplicity, we assume that the copolymer chain lengths are monodisperse, so that each chain contains an identical number $`L`$ of monomers. Even with this simplification, the problem is quite challenging because, as we shall see, a given parent can split into an arbitrarily large number of phases as the interaction strength is increased.
For this problem we choose as our polydisperse feature, $`\sigma `$, the difference between the fractions of A and B-monomers on a chain. This takes values in the range $`\sigma =1\mathrm{}1`$, with the extreme values $`\pm 1`$ corresponding to pure A and B chains, respectively. We now measure volumes in units of the chain volume (which is $`L`$ times larger than the monomeric volume used in the previous example). The total density $`\rho _0=๐\sigma \rho (\sigma )`$ is then just the copolymer volume fraction $`\varphi `$. (Note that for the case of length polydispersity, $`\varphi `$ was related to the first moment of the length distribution; here it is the zeroth moment of the distribution of chain compositions $`\sigma `$.) Within Flory-Huggins theory, the free energy density (in units of $`k_\text{B}T`$ per chain volume) is
$$f=๐\sigma \rho (\sigma )[\mathrm{ln}\rho (\sigma )1]+\frac{L}{L_\mathrm{s}}(1\rho _0)\mathrm{ln}(1\rho _0)+L\stackrel{~}{\chi }\rho _0(1\rho _0)L\stackrel{~}{\chi }^{}\rho _1^2$$
(92)
The interaction between A and B monomers enters through $`\rho _1=๐\sigma \sigma \rho (\sigma )`$. In (92), we have specialized to the case where the interactions of the A and B monomers with each other and with the solvent are symmetric in A and B; otherwise, there would be an additional term proportional to $`\rho _0\rho _1`$. We have also allowed the solvent to be a polymer of length $`L_\mathrm{s}`$; the case of a monomeric solvent is recovered for $`L_\mathrm{s}=1`$. In the following, we absorb factors of $`L`$ into the interaction parameters by setting
$`\chi =L\stackrel{~}{\chi },\chi ^{}=L\stackrel{~}{\chi }^{}`$
Defining also $`r=L/L_\mathrm{s}`$, the excess free energy takes the form
$$\stackrel{~}{f}=r(1\rho _0)\mathrm{ln}(1\rho _0)+\chi \rho _0(1\rho _0)\chi ^{}\rho _1^2.$$
(93)
The moment free energy (40), with only $`\rho _0`$ and $`\rho _1`$ retained, is then
$$f_\mathrm{m}=\lambda _0\rho _0+\lambda _1\rho _1\rho _0+r(1\rho _0)\mathrm{ln}(1\rho _0)+\chi \rho _0(1\rho _0)\chi ^{}\rho _1^2.$$
(94)
If extra moment densities with weight functions $`w_i(\sigma )`$ are included, these simply add a term $`\lambda _i\rho _i`$ each.
#### 1 Copolymer without solvent
We first consider the special case of a copolymer melt, for which the overall density (i.e., copolymer volume fraction) is constrained to take everywhere its maximum value $`\rho _0=1`$. The free energy (94) then simplifies to
$$f_\mathrm{m}=\lambda _0+\lambda _1\rho _1\chi ^{}\rho _1^2$$
(95)
up to irrelevant constants . Note that the same expression for the excess free energy of the dense system is obtained for a model in which the interaction energy between two species $`\sigma ,\sigma ^{}`$ varies as $`(\sigma \sigma ^{})^2`$:
$$\stackrel{~}{f}=\frac{1}{2}\chi ^{}๐\sigma ๐\sigma ^{}\rho (\sigma )\rho (\sigma ^{})(\sigma \sigma ^{})^2=\chi ^{}\rho _1^2+๐\sigma \rho (\sigma )\sigma ^2$$
(96)
The second term on the r.h.s. is linear in $`\rho (\sigma )`$ and can be discarded for the purposes of phase equilibrium calculations. The model (95) can therefore also be viewed as a simplified model of chemical fractionation, with $`\sigma `$ being related to aromaticity, for example, or another smoothly varying property of the different polymers. Eq. (96) suggests that the model should show fractionation into an ever-increasing number of phases as $`\chi ^{}`$ is increased: only the entropy of mixing prevents each value of $`\sigma `$ from forming a phase by itself, which would minimize this form of $`\stackrel{~}{f}`$. It is therefore an interesting test case for the moment free energy approach (and the method of adding further moment densities), yet simple enough for exact phase equilibrium calculations to remain feasible , allowing detailed comparisons to be made.
As a concrete example, we now consider phase separation from parent distributions of the form $`\rho ^{(0)}(\sigma )\mathrm{exp}(\gamma \sigma )`$ (for $`1\sigma 1`$, otherwise zero). The shape parameter $`\gamma `$ is then a fixed function of the parental first moment density $`\rho _1^{(0)}=๐\sigma \sigma \rho ^{(0)}(\sigma )`$. Fig 9 shows the exact coexistence curve for $`\rho _1^{(0)}=0.2`$, along with the predictions from our moment free energy with $`n`$ moment densities ($`\rho _i=๐\sigma \sigma ^i\rho (\sigma )`$, $`i=1\mathrm{}n`$) retained. Comparable results are found for other $`\rho _1^{(0)}`$. Even for the minimal set of moment densities ($`n=1`$) the point where phase separation first occurs on increasing $`\chi ^{}`$ is predicted correctly (this is the cloud point for the given parent). As more moments are added, the calculated coexistence curves approach the exact one to higher and higher precision. As expected, the precision decreases at high $`\chi ^{}`$, where fractionated phases proliferate; in this region, the number of coexisting phases predicted by the moment method increases with $`n`$. However, it is not always equal to $`n+1`$, as one might expect from a naive use of Gibbsโ phase rule; three-phase coexistence, for example, is first predicted for $`n=4`$. In fact, for fractionation problems such as this (but not more generally), study of the low temperature limit (large $`\chi ^{}`$) suggests that to obtain $`n+1`$ phases one may have to include up to $`2n`$ moments. We show below that using localized weight functions (rather than powers of $`\sigma `$) for the extra moments can reduce this number, but also gives less accurate coexistence curves.
In Fig. 10, we show how, for a given value of $`\chi ^{}`$, the lever rule is satisfied more and more accurately as the number of moment densities $`n`$ in the moment free energy is increased: the total density distribution $`\rho _{\mathrm{tot}}(\sigma )=_\alpha v^{(\alpha )}\rho ^{(\alpha )}(\sigma )`$ approaches the parent $`\rho ^{(0)}(\sigma )`$. This is by design, because the moment method forces the moment densities $`\rho _i=๐\sigma w_i(\sigma )\rho _{\mathrm{tot}}(\sigma )`$ of $`\rho _{\mathrm{tot}}(\sigma )`$ to be equal to those of the parent, for $`i=1\mathrm{}n`$; as $`n`$ is increased, therefore, $`\rho _{\mathrm{tot}}(\sigma )`$ approaches $`\rho ^{(0)}(\sigma )`$.
The above results for the dense copolymer model raise two general questions. First, how large does $`n`$ (the number of moment densities retained) have to be for the moment method to give reliable results for the coexisting phases (and, in particular, the correct number of phases)? And second, how should the weight functions for the extra moment densities be chosen? Regarding the first question, note first that the theoretical framework developed so far only says that the exact results (and therefore the correct number of phases) will be approached as $`n`$ gets large; it does not say for which value of $`n`$ a reasonable approximation is obtained. In fact it is clear that no โuniversalโ (problem-independent) value of $`n`$ can exist. The model studied in this section already provides a counterexample: As $`\chi ^{}`$ is increased, the (exact) number $`p`$ of coexisting phases increases without limit; therefore the minimum value of $`n`$ ($`n=p1`$) required to obtain this correct number of phases from a moment free energy calculation can also be arbitrarily large.
Nevertheless, we see already from Fig. 9 that, for the current example, the robustness of the results to the addition of extra moment densities provides a reasonable qualitative check of the accuracy of the coexistence curves. To study the convergence of the moment free energy results to the exact ones more quantitatively, especially in cases where the latter are not available for comparison, we need a measure of the deviation between $`\rho _{\mathrm{tot}}(\sigma )`$ and $`\rho ^{(0)}(\sigma )`$. To obtain a dimensionless quantity that does not scale with the overall density of the parent, we consider the normalized distributions $`n_{\mathrm{tot}}(\sigma )=\rho _{\mathrm{tot}}(\sigma )/\rho _{\mathrm{tot}}`$ and $`n^{(0)}(\sigma )=\rho ^{(0)}(\sigma )/\rho ^{(0)}`$ and define as our measure of deviation the โlog-errorโ
$$\delta =๐\sigma n^{(0)}(\sigma )\left(\mathrm{ln}\frac{n^{(0)}(\sigma )}{n_{\mathrm{tot}}(\sigma )}\right)^2$$
(97)
This becomes zero only when $`n_{\mathrm{tot}}(\sigma )=n^{(0)}(\sigma )`$; otherwise it is positive. When the deviations between the two distributions are small, the logarithm becomes to leading order the relative deviation $`n_{\mathrm{tot}}(\sigma )/n^{(0)}(\sigma )1`$ and we can identify $`\delta ^{1/2}`$ as the root-mean-square relative deviation. We work with the logarithm rather than directly with the relative deviation because the former gives more sensible behavior for larger deviations; in particular, isolated points where $`n_{\mathrm{tot}}(\sigma )`$ is nonzero while $`n^{(0)}(\sigma )`$ is close to zero do not lead to divergences of $`\delta `$ .
Fig. 11 shows the $`n`$-dependence of the log-error $`\delta `$ for three fixed values of $`\chi ^{}`$, with the same parent as in Fig 9 . The correct number of phases (3, 4, 5, respectively) is reached at values of $`\delta 4\times 10^3`$, $`5\times 10^4`$, $`2\times 10^3`$ for $`\chi =5`$, 10, 15, respectively. This suggests that the log-error $`\delta `$ might generally be a useful heuristic criterion for guiding the choice of $`n`$, with values of $`\delta `$ around $`10^4`$ (corresponding to an average deviation between $`n_{\mathrm{tot}}(\sigma )`$ and $`n^{(0)}(\sigma )`$ of around 1%) ensuring that the correct number of phases has been detected. Note that, although for the current example we did know the correct number of phases (because the simplicity of the problem made an exact calculation feasible), the definition of $`\delta `$ does not require us to know in advance any properties of the exact solution of the coexistence problem. Also, note that in real physical materials it is extremely unlikely for the parent or โfeedstockโ distribution $`\rho ^{(0)}(\sigma )`$ to be known to better than 1% accuracy. Hence even for systems, such as polymers, where a mean-field (and thus truncatable) free energy model is thought to be highly accurate, deviations from the lever rule, at about the 1% level, will almost never be the main source of uncertainty in phase diagram prediction.
We now turn to the second general question, regarding the choice of weight functions for the extra moment densities. Comparing (63) with the formally exact solution (62) of the coexistence problem tells us at least in principle what is required: the log-ratio $`\mathrm{ln}\stackrel{~}{R}(\sigma )/\rho ^{(0)}(\sigma )`$ of the โeffective priorโ and the parent needs to be well approximated by a linear combination of the weight functions of the extra moment densities. However, the effective prior is unknown (otherwise we would already have the exact solution of the phase coexistence problem), and so this criterion is of little use .
To make progress, let us try to gain some intuition from the example at hand (dense random copolymers). So far, we have simply taken increasing powers of $`\sigma `$ for the extra weight functions. On the other hand, looking at Fig. 10 one might suspect that localized weight functions might be more suitable for capturing the fractionating of the parent into the various daughter phases. We therefore tried the simple โtriangleโ and โbellโ weight functions shown in Fig. 12(a). For a given $`n`$, there are $`n1`$ extra moment densities (the first moment density, with weight function $`w_1(\sigma )=\sigma `$, is prescribed by the excess free energy) and we take these to be evenly spaced across the range $`\sigma =1\mathrm{}1`$, with overlaps as indicated in Fig 12(a). (For the case of triangular weight functions, the log ratio $`\mathrm{ln}\rho (\sigma )/\rho ^{(0)}(\sigma )=_i\lambda _iw_i(\sigma )`$ between the density distribution of any phase and that of the parent is thus approximated by a โtrapezoidalโ function across $`n`$ equal intervals spanning $`1\mathrm{}1`$.) Fig. 12(b) shows the corresponding coexistence curves for the triangle case with $`n=1,2,3,4`$. Note that multi-phase coexistence is generally detected for smaller $`n`$ than for the case of power-law weight functions (Fig. 9), in agreement with the expectations outlined above. However, the predicted number of phases does not vary monotonically with $`n`$: in the range of $`\chi ^{}`$ shown in Fig. 12(b), up to four phases are predicted for $`n=3`$, whereas $`n=4`$ gives no more than three. This effect illustrates a significant difference between power-law and triangle weight functions. The former, but not the latter, make up an incremental set: increasing $`n`$ by one corresponds to adding a new weight function while leaving the existing set of weight functions unchanged. The family (42) of accessible density distributions $`\rho (\sigma )`$ is thus enlarged without losing any of its former members. Intuitively, one then expects that the number of phases cannot decrease as $`n`$ is increased. But for nonincremental sets, such as triangle weight functions (and bells, which give qualitatively similar results to triangles), each increase in $`n`$ creates a larger family but completely discards the previous one, so that there is then no reason for the results to be monotonic in $`n`$.
We regard the fact that the number of phases can go up and down with increasing $`n`$ as a disadvantage of non-incremental weight functions. In the current example, we also find that they generally give higher values of the log-error than power-law weight functions; see Fig. 12(c). Finally, incremental sets of weight functions also have the benefit that the results for $`n`$ can be used to initialize the calculation for $`n+1`$, thus helping to speed up the numerics (though we have not yet exploited this property). On the basis of these observations, we would generally recommend the use of incremental sets of weight functions for extra moment densities. Beyond this, the optimal choice of extra weight functions remains largely an open problem. One avenue worth exploring in future work might be the possibility of choosing extra weight functions adaptively. One could imagine monitoring the log-error in a phase coexistence calculation, and increasing the number of extra moments by one every time it becomes larger than a certain threshold ($`10^4`$, say). The new weight function could then be chosen as a best fit to the current lever-rule violation $`\mathrm{ln}\rho _{\mathrm{tot}}(\sigma )/\rho ^{(0)}(\sigma )`$ (for example as a linear combination of weight functions from some predefined โpoolโ). We are planning to investigate this method in future work.
To conclude this section on the dense random copolymer model, we briefly discuss the spinodal criterion, and ask whether critical or multi-critical points can exist for a general parent distribution $`\rho ^{(0)}(\sigma )`$ (with $`\rho _0^{(0)}=1`$). The criterion (56), applied to our one-moment free energy $`f_\mathrm{m}(\rho _1)`$ becomes simply
$$f_\mathrm{m}(\rho _1)=๐ช\left((\rho _1\rho _1^{(0)})^{l+1}\right)$$
(98)
where $`l=2`$ for a spinodal and $`l=2n1`$ for an $`n`$-critical point. From eq. (95), we have that
$$f_\mathrm{m}=\lambda _0+\lambda _1\rho _1\chi ^{}\rho _1^2=s_\mathrm{m}\chi ^{}\rho _1^2$$
(99)
The excess part $`\chi ^{}\rho _1^2`$ is quadratic in $`\rho _1`$ and therefore only enters the spinodal criterion; the additional conditions for critical points only depend on the moment entropy $`s_\mathrm{m}=\lambda _0\lambda _1\rho _1`$. Because of the dense limit constraint $`\rho _1=1`$, $`\lambda _0`$ is related to $`\lambda _1`$ by
$`\lambda _0=\mathrm{ln}{\displaystyle ๐\sigma \rho ^{(0)}(\sigma )e^{\lambda _1\sigma }}h(\lambda _1)`$
As expected from the general discussion in Sec. II C, $`\lambda _0`$ is thus the cumulant generating function $`h(\lambda _1)`$ of $`\rho ^{(0)}(\sigma )`$ ($`n^{(0)}(\sigma )`$ in the dense limit we are considering), introduced in eq. (36) within the combinatorial approach. The moment entropy $`s_\mathrm{m}=h\lambda _1\rho _1`$ is its Legendre transform; its behavior for $`\rho _1\rho _1^{(0)}`$ reflects that of $`h`$ for small $`\lambda _1`$. We can thus use the expansion $`h=_{j=1}^{\mathrm{}}(c_j/j!)\lambda _1^j`$, where the $`c_j`$ are the cumulants of $`\rho ^{(0)}(\sigma )`$; $`c_1\rho _1^{(0)}`$ is its mean, $`c_2`$ its variance etc.. Using the properties of the Legendre transform, it then follows that the leading term of the moment entropy is $`s_\mathrm{m}=(\rho _1\rho _1^{(0)})^2/(2c_2)`$. Inserting this into (99), we see immediately that the spinodal condition becomes
$$2\chi ^{}c_2=2\chi ^{}\left(\rho _2^{(0)}(\rho _1^{(0)})^2\right)=1$$
(100)
To find the additional conditions for critical points, consider what happens if the cumulants $`c_3`$, โฆ, $`c_l`$ are zero, hence $`h=\rho _1^{(0)}\lambda _1+c_2\lambda _1^2/2+๐ช(\lambda _1^{l+1})`$. It can then easily be shown that the moment entropy behaves as $`s_\mathrm{m}=(\rho _1\rho _1^{(0)})^2/(2c_2)+๐ช((\rho _1\rho _1^{(0)})^{l+1})`$; conversely, one can show that this behavior of $`s_\mathrm{m}`$ implies that the cumulants $`c_3`$, โฆ, $`c_l`$ are zero. We thus have the simple condition that for an $`n`$-critical point to be observed, the cumulants $`c_3`$, $`\mathrm{}`$, $`c_{2n1}`$ of the parent distribution have to be zero; the value of $`\chi ^{}`$ at this critical point is given by the spinodal condition (100). (For an ordinary critical point, $`n=2`$, the condition $`c_3=0`$ is well known in this context .) We therefore come to the somewhat surprising conclusion that even in a very simple polydisperse system such as this โ with a single moment density occurring in the excess free energy โ multi-critical points of arbitrary order can occur, at least in principle. In practice, the required fine-tuning of the parent will of course make experimental study of these points difficult.
#### 2 Copolymer with solvent
We now consider the random copolymer model in the presence of solvent, i.e., for a copolymer volume fraction $`\rho _1<1`$. We are not aware of previous work on this model in the literature, but will briefly discuss below the link to models of homopolymer/copolymer mixtures . The excess free energy (93) then depends on two moment densities, rather than just one as in all previous examples. For simplicity, we restrict ourselves to the case of a neutral solvent which does not in itself induce phase separation; this corresponds to $`\chi =0`$, making the excess free energy
$$\stackrel{~}{f}=r(1\rho _0)\mathrm{ln}(1\rho _0)\chi ^{}\rho _1^2$$
(101)
This is still sufficiently simple that the exact spinodal condition (58) โ for a system with a general density distribution $`\rho (\sigma )`$ โ can be worked out analytically; one finds
$$2\chi ^{}=\frac{r\rho _0+1\rho _0}{r(\rho _2\rho _0\rho _1^2)+\rho _2(1\rho _0)}$$
(102)
In the dense limit $`\rho _01`$, the earlier result (100) is recovered. To study critical and multi-critical points, we use the moment free energy corresponding to (101),
$$f_\mathrm{m}=\lambda _0\rho _0+\lambda _1\rho _1\rho _0+r(1\rho _0)\mathrm{ln}(1\rho _0)\chi ^{}\rho _1^2$$
(103)
and start from the criterion (56). The moment chemical potentials are given by
$`\mu _0`$ $`=`$ $`\lambda _0r[\mathrm{ln}(1\rho _0)1]`$ (104)
$`\mu _1`$ $`=`$ $`\lambda _12\chi ^{}\rho _1`$ (105)
and the curve $`๐(ฯต)`$ referred to in (56) โ along which the critical phases merge โ is simply parameterized in terms of $`\lambda _0(ฯต)`$ and $`\lambda _1(ฯต)`$. The restriction that the curve must pass through the parent at $`ฯต=0`$ means that $`\lambda _0(0)=\lambda _1(0)=0`$, so to get the small $`ฯต`$ behavior of the chemical potentials (105) we can expand for small $`\lambda _0`$ and $`\lambda _1`$. To simplify the algebra, we restrict ourselves to parent distributions $`\rho ^{(0)}(\sigma )`$ which are symmetric about $`\sigma =0`$. Using this symmetry and the fact that $`ฯต`$ is simply a dummy parameter which can be reparameterized arbitrarily, we can then set $`\lambda _1=ฯต`$ and $`\lambda _0=a_2ฯต^2+a_4ฯต^4+๐ช(ฯต^6)`$.
To have a spinodal, the moment chemical potentials (105) must differ from those of the parent by no more than $`๐ช(ฯต^2)`$. We find that this gives the condition $`2\chi ^{}=1/\rho _2^{(0)}`$, in agreement with (102): symmetric parents have $`\rho _1^{(0)}=0`$. For a critical point, the chemical potential difference must be decreased to $`๐ช(ฯต^3)`$. This gives no new condition (but determines $`a_2`$), implying that parents with symmetric distributions are automatically critical at their spinodal. This result has its origin in the symmetry $`\sigma \sigma `$ of the excess free energy, and can be understood by considering the $`\rho _0`$, $`\rho _1`$, $`\chi ^{}`$ phase diagram of the moment free energy (103). In this phase diagram there is a two-dimensional spinodal surface, and within this surface lies a line of critical points. The symmetry $`\sigma \sigma `$ (corresponding to $`\rho _1\rho _1`$, $`\lambda _1\lambda _1`$) forces this critical line to lie within the plane $`\rho _1=0`$ of symmetrical distributions, implying that for such distributions spinodal and critical points coincide.
For a tricritical point, finally, where the chemical potential difference must be further reduced to $`๐ช(ฯต^3)`$, we find after some algebra the condition
$`\rho _4^{(0)}={\displaystyle \frac{3r(\rho _2^{(0)})^2}{r\rho _0^{(0)}+1\rho _0^{(0)}}}`$
As an example, consider a uniform parent distribution $`\rho ^{(0)}(\sigma )=\text{const}`$, which can be written as $`\rho ^{(0)}(\sigma )=\rho _0^{(0)}/2`$, using the fact that $`\rho _0^{(0)}=_1^1๐\sigma \rho ^{(0)}(\sigma )`$. Then $`\rho _2^{(0)}=\rho _0^{(0)}/3`$ and $`\rho _4^{(0)}=\rho _0^{(0)}/5`$ and so a tricritical point occurs if the overall density (i.e., the copolymer volume fraction) is $`\rho _0^{(0)}=3/(2r+3)`$. Fig. 13 shows the coexistence curve calculated for this parent (with $`r=1`$), which clearly shows the tricritical point at the predicted value $`\chi ^{}=1/(2\rho _2^{(0)})=r+3/2=2.5`$. Our numerical implementation manages to locate the tricritical point and follow the three coexisting phases without problems; we take that as a signature of its robustness . Note that the tricritical point that we found is closely analogous to that studied by Leibler for a symmetric blend of two homopolymers and a symmetric random copolymer that is, nonetheless, chemically monodisperse (in the sense that $`\sigma =0`$ for all copolymers present). In fact, in our notation, the scenario of Ref. simply corresponds to a parent density of the form $`\rho ^{(0)}(\sigma )\delta (\sigma 1)+\delta (\sigma +1)`$, with the copolymer ($`\sigma =0`$) now playing the role of the neutral solvent.
We conclude this section with another illustration of the geometrical intuition which the moment free energy can provide. In Fig. 14, we plot the moment free energy $`f_\mathrm{m}(\rho _0,\rho _1)`$, eq. (103), for the uniform parent $`\rho ^{(0)}(\sigma )`$ studied above. Two values of $`\chi ^{}`$ are shown: the tricritical value ($`\chi ^{}=2.5`$), and a value just in the three-phase region ($`\chi ^{}=2.65`$); the developing three minima of the free energy can clearly be seen. Even though the underlying polydisperse system has an infinite number of degrees of freedom $`\rho (\sigma )`$, the moment method thus gives us a tool for understanding and visualizing the occurrence of the tricritical point in terms of a simple free energy surface with two independent variables.
## VI Conclusion
Polydisperse systems contain particles with an attribute $`\sigma `$ which takes values in a continuous range (such as particle size in colloids or chain length in polymers). They therefore have an infinite number of conserved densities, corresponding to a density distribution $`\rho (\sigma )`$. The fact that the free energy can depend on all details of $`\rho (\sigma )`$ makes the analysis of the phase behavior of such systems highly non-trivial. However, in many (especially mean-field) models the excess free energy only depends on a finite number of moment densities, i.e., (generalized) moments of $`\rho (\sigma )`$; the only dependence on other details of $`\rho (\sigma )`$ is through the ideal part of the free energy. For such models, which we call truncatable, we showed how a moment free energy can be constructed, which only depends on the moment densities appearing in the excess free energy. Two possible approaches for the construction of the moment free energy were described in Sec. II. The first, which we call the projection method (Sec. II A), is primarily geometrical; the second is mainly combinatorial (Sec. II B) and starts from first principles to derive an expression for the ideal part of the free energy in terms of moment densities. In Sec. II C, we showed that these two approaches give essentially equivalent results.
In Sec. III, we explored the properties of the moment free energy in detail. The moment free energy depends only on a finite number of moment densities, and can be used to predict phase behavior using the conventional methods known from the thermodynamics of finite mixtures. We showed that, for any truncatable model, this procedure yields the same spinodals, critical and multi-critical points as the original free energy, even though the latter depends on all details of the density distribution $`\rho (\sigma )`$. It also correctly predicts the onset of phase coexistence, i.e., the cloud points and corresponding shadows. Beyond the onset of phase coexistence (i.e., when one or more coexisting phases occupy comparable volumes), the results from the moment free energy are approximate, but can be made arbitrarily accurate by retaining additional moment densities in the moment free energy. We also explained how to check the local and global stability of a phase coexistence calculated from the moment free energy, and how the moment free energy reflects overall phase behavior in the space of density distributions $`\rho (\sigma )`$.
We discussed aspects of the numerical implementation of the moment free energy method in Sec. IV, highlighting its many advantages in terms of robustness, efficiency and stability. The method allows violations of the lever rule, but then ensures that these be kept small by adding $`nK`$ extra moments (where $`K`$ is the number of moment densities actually present in the excess free energy of the underlying truncatable model). The solution of the phase equilibrium conditions thus proceeds in an $`n`$-dimensional space; but global (tangent plane) stability can always be verified by minimimization (of the tangent plane distance) in a space of dimension $`K`$. For $`K>1`$ exact minimization is not practicable, but standard search methods can instead be employed; we chose a Monte-Carlo type algorithm for this.
We then surveyed several sample applications in Sec. V. For length-polydisperse homopolymers (treated within Flory-Huggins theory), we showed explicitly how the moment free energy with only a single moment density (the volume fraction of polymer) gives the correct spinodals and critical points. By retaining the chain number density as an additional moment density, we could rationalize the behavior of cloud point and shadow curves as cuts through conventional two-species phase diagrams. As a second example, the case of copolymers with chemical polydispersity was considered, both without and with solvent. This gave us the opportunity to test the performance of the moment free energy method for multi-phase coexistence, and in particular to study how the accuracy of the predictions in the coexistence region increases as more and more moment densities are retained. In the presence of solvent, the same theory predicts a tricritical point under appropriate conditions, which was detected without problems, along with the associated three phase region, by our numerical algorithm. A plot of the two-moment free energy gave a simple geometrical picture of the origin of this tricritical point.
The results for the copolymer model showed that the log-error (97) can be used to decide when enough moment densities have been included to locate the correct number of phases and their compositions. For general purposes we suggest a log-error tolerance of order $`10^4`$; this corresponds to limiting the rms lever-rule violations at the predicted phase coexistence (found by global minimization of the moment free energy) to 1% or so of the parent distribution. Given that in practice the parent is itself not known to such high precision, these violations lie well within an acceptable range of error. Indeed, every state of coexistence actually predicted by our algorithm represents the exact phase equilibrium of some or other parent, indistinguishable from the real one to within experimental accuracy. Here, as elsewhere in this paper, โexactโ means exactly as predicted by a truncatable model free energy; such models are, of course, themselves approximate, which is another reason not to expend resources reducing the log-error below the level suggested above.
In summary, with exact results for cloud point and shadow curves, critical points and spinodals, as well as refinably accurate coexistence curves and multi-phase regions, the moment free energy method allows rapid and accurate computation of the phase behavior of many polydisperse systems. Moreover, by establishing the link to a projected free energy $`f_\mathrm{m}(\rho _i)`$ which is a function of a finite set of conserved moment densities $`\rho _i`$, it restores to the problem much of the geometrical interpretation and insight (as well as the computational methodology) associated with the phase behavior of finite mixtures. This contrasts with many procedures previously in use for truncatable polydisperse systems . Some previous approximations to the problem have used (generalized) moments as thermodynamic coordinates; see e.g. . The moment free energy method provides a rational basis for these methods and, so long as the generalized moments are correctly chosen, guarantees that many properties of interest are found exactly. Note also that the commonplace method of โbinningโ the $`\sigma `$-distribution into discrete โpseudo-componentsโ can be seen as a special case of the moment method in which each weight function is zero outside the corresponding bin; but since these moments do not, in general, coincide with those contained in the excess free energy the advantages described above are then lost.
We finish with some remarks about extensions to the moment method, within and outside the sphere of phase diagram prediction. First, methods of this type may extend to nontruncatable models, for which the excess free energy cannot be written directly in terms of a finite number of moments as in (6). For example, many mean-field theories correspond to a variational minimization of the free energy: $`FE_0TS_0`$, where the subscript $`0`$ refers to a trial Hamiltonian . In such a case, one might choose to first make a physically motivated decision about which (and how many) moment densities $`\rho _i`$ to keep, and then include among the variational parameters the โtransverseโ degrees of freedom of the density distribution $`\rho (\sigma )`$. We have not pursued this approach, though it may form a promising basis for future developments. Secondly, we note that physical insight based on moment free energies may increasingly play a part in understanding kinetic problems. For example, one of us has argued that in many systems the zeroth moment (mean particle density) can relax more rapidly (by collective diffusion) than can the higher moments (by interdiffusion of different species), and that this may lead to novel kinetics in certain areas of the phase diagram. And finally, we point to recent work in which the moment method has been extended, via density functional theory, to allow study of inhomogeneous states; this opens the way to studying the effect of polydispersity on interfacial tensions and other interfacial thermodynamic properties.
Acknowledgements: We thank P. Bladon, N. Clarke, R. M. L. Evans, T. McLeish, P. Olmsted, I. Pagonabarraga, W. C. K. Poon and R. Sear for helpful discussions. PS is grateful to the Royal Society for financial support. This work was also funded in part under EPSRC GR/M29696 and GR/L81185.
## A Moment (Gibbs) free energy for fixed pressure
In this appendix, we explain how the construction of the moment free energy is extended to scenarios where the pressure (rather than the overall system volume) is held constant. One then describes the system by the particle number distribution $`N(\sigma )=V\rho (\sigma )`$; this is normalized so that its integral gives the total number of particles, $`N=๐\sigma N(\sigma )`$. The relevant thermodynamic potential is the Gibbs free energy $`G=\mathrm{min}_VF+\mathrm{\Pi }V`$, which we now construct. For a truncatable system, whose free energy density obeys eqs. (3,4) by definition, the Helmholtz free energy can be written as
$`F=T{\displaystyle ๐\sigma N(\sigma )\left[\mathrm{ln}\frac{N(\sigma )}{V}1\right]}+\stackrel{~}{F}`$
where the excess free energy $`\stackrel{~}{F}(N_i,V)=V\stackrel{~}{f}(\rho _i)`$ depends on volume $`V`$ and on the moments $`N_i=V\rho _i=๐\sigma N(\sigma )w_i(\sigma )`$ of the particle number distribution. If $`\stackrel{~}{F}=0`$, a Legendre transform gives the Gibbs free energy of an ideal mixture:
$`G_{\mathrm{id}}=\underset{V}{\mathrm{min}}F_{\mathrm{id}}+\mathrm{\Pi }V=T{\displaystyle ๐\sigma N(\sigma )\mathrm{ln}\frac{\mathrm{\Pi }N(\sigma )}{NT}}=NT{\displaystyle ๐\sigma n(\sigma )\mathrm{ln}n(\sigma )}+NT\mathrm{ln}{\displaystyle \frac{\mathrm{\Pi }}{T}}`$
Here $`n(\sigma )=N(\sigma )/N=\rho (\sigma )/\rho `$ is the normalized particle distribution. In the general case, we split off the ideal part
$`G=G_{\mathrm{id}}+\stackrel{~}{G}=\underset{V}{\mathrm{min}}F+\mathrm{\Pi }V`$
and obtain for the excess Gibbs free energy
$`\stackrel{~}{G}=\underset{V}{\mathrm{min}}NT\mathrm{ln}{\displaystyle \frac{NT}{\mathrm{\Pi }V}}+\stackrel{~}{F}(N_i,V)+\mathrm{\Pi }V`$
This is a function of $`N`$ and the $`N_i`$ (as well as the fixed control parameters $`\mathrm{\Pi },T`$, which we suppress in our notation). We can therefore write the excess Gibbs free energy per particle, $`\stackrel{~}{G}(N_i,N)/N=\stackrel{~}{g}(m_i)`$, as a function of the moments $`m_i=N_i/N`$ ($`i=1\mathrm{}K`$) of the normalized particle distribution $`n(\sigma )`$. Altogether, the Gibbs free energy per particle of a truncatable system becomes
$$g[n(\sigma )]=G/N=T๐\sigma n(\sigma )\mathrm{ln}n(\sigma )+T\mathrm{ln}\frac{\mathrm{\Pi }}{T}+\stackrel{~}{g}(m_i)$$
(A1)
This has again a truncatable structure; the excess part $`\stackrel{~}{g}`$ โinheritsโ its moment structure from $`\stackrel{~}{f}`$. Note that because of the normalization of the $`m_i`$, $`\stackrel{~}{g}`$ does not depend on the density $`\rho \rho _0`$. It is therefore normally a function of one less variable than $`\stackrel{~}{f}`$, unless $`\stackrel{~}{f}`$ is already independent of $`\rho `$ (as is the case in Flory-Huggins theory, eq. (71), for example). The chemical potentials $`\mu (\sigma )`$ follow from (A1) as
$`\mu (\sigma )`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta N(\sigma )}}Ng[N(\sigma )/N]`$ (A2)
$`=`$ $`{\displaystyle \frac{\delta g}{\delta n(\sigma )}}+g{\displaystyle ๐\sigma ^{}n(\sigma ^{})\frac{\delta g}{\delta n(\sigma ^{})}}`$ (A3)
$`=`$ $`T\mathrm{ln}n(\sigma )+{\displaystyle \underset{i}{}}w_i(\sigma ){\displaystyle \frac{\stackrel{~}{g}}{m_i}}+\left[T\mathrm{ln}{\displaystyle \frac{\mathrm{\Pi }}{T}}+\stackrel{~}{g}{\displaystyle \underset{i}{}}m_i{\displaystyle \frac{\stackrel{~}{g}}{m_i}}\right]`$ (A4)
Multiplying by $`n(\sigma )`$ and integrating over $`\sigma `$, one has $`g=๐\sigma \mu (\sigma )n(\sigma )`$ as it should be (compare eq. (A1)). Comparing (A4) with (48), one deduces that
$`{\displaystyle \frac{\stackrel{~}{g}}{m_i}}=\stackrel{~}{\mu }_i`$
and that the density $`\rho `$ of a phase with moments $`m_i`$ is given by
$`T\mathrm{ln}\left({\displaystyle \frac{T\rho }{\mathrm{\Pi }}}\right)=\stackrel{~}{g}{\displaystyle \underset{i}{}}m_i\stackrel{~}{\mu }_i.`$
To construct the moment Gibbs free energy, one can now proceed as in the constant volume case: in the ideal contribution, we replace $`\mathrm{ln}n(\sigma )\mathrm{ln}[n(\sigma )/n^{(0)}(\sigma )]`$, with a normalized parent distribution $`n^{(0)}(\sigma )`$, and then minimize $`g`$ at fixed values of the $`m_i`$. The minimum occurs for distributions $`n(\sigma )`$ from the family
$$n(\sigma )=n^{(0)}(\sigma )\mathrm{exp}\left(\lambda _0+\underset{i}{}\lambda _iw_i(\sigma )\right)$$
(A5)
The extra Lagrange multiplier $`\lambda _0`$ comes from the normalization condition $`m_0=๐\sigma n(\sigma )=1`$. Inserting into (A1), one obtains the moment Gibbs free energy
$$g_\mathrm{m}(m_i)=T\left(\lambda _0+\underset{i}{}\lambda _im_i\right)+T\mathrm{ln}\frac{\mathrm{\Pi }}{T}+\stackrel{~}{g}(m_i)$$
(A6)
where the $`m_i`$ are given by
$`m_i={\displaystyle ๐\sigma w_i(\sigma )n^{(0)}(\sigma )\mathrm{exp}\left(\lambda _0+\underset{i}{}\lambda _iw_i(\sigma )\right)}={\displaystyle \frac{๐\sigma w_i(\sigma )n^{(0)}(\sigma )\mathrm{exp}\left(_i\lambda _iw_i(\sigma )\right)}{๐\sigma n^{(0)}(\sigma )\mathrm{exp}\left(_i\lambda _iw_i(\sigma )\right)}}`$
The derivatives of $`g_\mathrm{m}`$ can again be identified as moment chemical potentials and obey
$$\mu _i\frac{g_\mathrm{m}}{m_i}=T\lambda _i+\frac{\stackrel{~}{g}}{m_i}=T\lambda _i+\stackrel{~}{\mu }_i.$$
(A7)
The constraint $`m_0=1`$ does not affect this result because the partial derivative is taken with the values of all moments other than $`m_i`$ fixed anyway. For the same reason, the second derivatives of the ideal part of $`g_\mathrm{m}`$ are (up to a factor $`T`$) given by the inverse of the matrix of second-order normalized moments $`m_{ij}`$, by analogy with (46). (Note that the first row and column of the second-order moment matrix, corresponding to $`i=0`$ or $`j=0`$, need to be retained; they can only be discarded after the inverse has been taken.)
From (A4), one deduces immediately that if one of a number of coexisting phases is in the family (A5), then so are all others. Equality of the exact chemical potentials $`\mu (\sigma )`$ in all phases is then equivalent to equality of the quantities
$$\stackrel{~}{\mu }_i+T\lambda _i=\mu _i(i=1\mathrm{}K),T\lambda _0+\stackrel{~}{g}\underset{i}{}m_i\stackrel{~}{\mu }_i.$$
(A8)
But from (A6,A7), these are simply the moment chemical potentials $`\mu _i=g_\mathrm{m}/m_i`$ and (up to an unimportant additive constant) the associated Legendre transform $`g_\mathrm{m}_im_i\mu _i`$. Therefore, phase equilibria can be constructed by applying the usual tangent plane construction to the moment Gibbs free energy, in analogy with the constant volume case. Note, however, that this tangent plane construction now takes place in the space of the normalized moments $`m_i`$, rather than that of the unnormalized moment densities $`\rho _i`$. As pointed out above, this generically reduces the dimension by one. For the trivial case of a monodisperse system, $`\stackrel{~}{f}`$ is a function of $`\rho `$ only and $`\stackrel{~}{g}`$ does not depend on any density variables at all, so that the tangent plane condition becomes void as expected.
The exactness statements in Sec. III can also be directly translated to the constant pressure case. The arguments above imply directly that the onset of phase coexistence is found exactly from the moment Gibbs free energy: all phases are in the family (A5), because one of them (the parent) is, and the requirement of equal chemical potentials is satisfied. Spinodals and (multi-) critical points are also found exactly. Arguing as in Sec. III A, and using the vector notation of (56), the criterion for such points is found as
$$\mathrm{\Delta }๐๐(๐(ฯต))๐(๐^{(0)})=๐ช(ฯต^l)$$
(A9)
where $`l=2`$ for a spinodal and $`l=2n1`$ for an $`n`$-critical point. Again, the curve $`๐(ฯต)`$ only has to pass through phases with equal chemical potentials and can therefore be chosen to lie within the family (A5). As discussed above (eq. (A8)), the chemical potential difference on the left hand side of (A9) is then zero to the required order in $`ฯต`$ if the same holds for the $`\mathrm{\Delta }\mu _i`$ and $`\mathrm{\Delta }h`$, where $`h=g_\mathrm{m}_im_i\mu _i`$. It may look as if the constraint on $`h`$ gives one more condition here than in the constant volume case; but in fact $`\mathrm{\Delta }\mu _i=๐ช(ฯต^l)`$ for all $`i`$ already implies that $`\mathrm{\Delta }g_\mathrm{m}=๐ช(ฯต^{l+1})`$ and hence $`\mathrm{\Delta }h=๐ช(ฯต^l)`$. The exact condition (A9) for spinodals and critical points is therefore again equivalent to the same condition obtained from the moment Gibbs free energy.
## B Moment entropy of mixing and large deviation theory
In this appendix, we discuss some interesting properties of the Legendre transform result (31) for the moment entropy of mixing, in particular its relation to large deviation theory (LDT).
We first note that (31), which we derived by taking the limit $`x0`$ of the result (28) for general $`x`$, can also be obtained by a more direct route. In the limit $`x0`$, the sizes of particles in the smaller system become independent random variables drawn from $`n^{(0)}(\sigma )`$; the second phase can be viewed as a reservoir to which the small phase is connected. One writes the moment generating function for $`๐ซ(m)`$ in the small phase as a product of $`xN`$ independent moment generating functions of $`n^{(0)}(\sigma )`$, and evaluates the integral over $`๐ซ(m)`$ by a saddle point method .
One can also view (31) as a generalisation of Cramรฉrโs large deviation theory (LDT) . Recall that this theory treats the โwingsโ of a distribution like $`๐ซ(m)`$ correctly, to which the central limit theorem (CLT) does not apply . However, the CLT approximation in this problem has quite a beautiful interpretation, and is worth describing separately. It can be derived by assuming that $`n^{(0)}(\sigma )`$ is itself a Gaussian in eq. (31), or just written down directly:
$`๐ซ(m)\mathrm{exp}\left[N{\displaystyle \frac{(mm^{(0)})^2}{2v^{(0)}}}\right]`$
where $`v^{(0)}`$ is the variance of the parent distribution. We see that this gives a term in the free energy of the form
$`\mathrm{\Delta }f=T\rho s=\rho T{\displaystyle \frac{(mm^{(0)})^2}{2v^{(0)}}}.`$
This result states that there is an entropic spring term in the free energy that penalizes deviations of the mean size $`m`$ from the mean size in the parent distribution. The spring constant is inversely proportional to the variance in the parent distribution, thus if the parent is narrower, it is harder to move away from the parental mean size . However, as indicated above, the CLT is not sufficiently accurate for our purposes. For instance it gives a finite weight to $`m<0`$ even if $`w(\sigma )`$ is strictly positive. That said, the CLT may be attractive for narrow size distributions, and one might make a connection with the recent results of Evans et al (even though the latter do not seem to rely on a specific shape of the distribution). We have not fully explored this avenue.
## C Spinodal criterion from exact free energy
In this appendix, we apply the spinodal criterion (53) to the exact free energy (41) and show that it can be expressed in a form identical to (58). This result has been given by a number of authors , but we include it here for the sake of completeness.
Choosing for $`๐`$ the vector $`\rho (\sigma )`$ (whose โcomponentsโ are the values of $`\rho (\sigma )`$ for all $`\sigma `$), the spinodal criterion (53) applied to (41) becomes
$$๐\sigma ^{}\frac{\delta ^2f}{\delta \rho (\sigma )\delta \rho (\sigma ^{})}\delta \rho (\sigma ^{})=\underset{i,j}{}\frac{^2\stackrel{~}{f}}{\rho _i\rho _j}w_i(\sigma )\delta \rho _j+T\frac{\delta \rho (\sigma )}{\rho (\sigma )}=0$$
(C1)
The change of $`\mathrm{ln}\rho (\sigma )`$ along the instability direction (which is $`\delta \rho (\sigma )/\rho (\sigma )`$) is therefore a linear combination of the weight functions $`w_i(\sigma )`$. This means that the instability direction is within the family (42), consistent with our general discussion in Sec. III A. One can now rewrite (C1) in the form
$$\delta \rho (\sigma )=\beta \underset{i,j}{}w_i(\sigma )\rho (\sigma )\frac{^2\stackrel{~}{f}}{\rho _i\rho _j}\delta \rho _j$$
(C2)
and take the moment with the $`k`$-th weight function $`w_k(\sigma )`$ to get
$`\delta \rho _k+\beta {\displaystyle \underset{i,j}{}}\rho _{ki}{\displaystyle \frac{^2\stackrel{~}{f}}{\rho _i\rho _j}}\delta \rho _j=0.`$
As promised, this is identical to the spinodal condition (58) derived from the moment free energy, with the matrix multiplications written out explicitly.
## D Determinant form of critical point criterion
In this appendix, we give the form of the critical point criterion (60) that uses the spinodal determinant $`Y`$ from (58. At a critical point, the instability direction connects two neighboring points on the spinodal. The first order variation of the spinodal determinant $`Y`$ along this direction must therefore vanish. This gives
$$\delta Y=\underset{i}{}\frac{Y}{\rho _i}\delta \rho _i+\underset{ij}{}\frac{Y}{\rho _{ij}}\delta \rho _{ij}=0.$$
(D1)
Here the $`\delta \rho _{ij}`$ are the changes in the second order moment densities along the instability direction. We can express these as a function of the $`\delta \rho _i`$ via the change in the Lagrange multipliers $`\lambda _i`$: From the definition (43), a change in the $`\lambda _i`$ changes the first and second order moment densities according to
$$\frac{\rho _i}{\lambda _j}=\rho _{ij}=๐ด_{ij},\frac{\rho _{ij}}{\lambda _k}=\rho _{ijk}.$$
(D2)
Hence changes in second and first order moment densities are related by
$`\delta \rho _{kl}={\displaystyle \underset{i}{}}\rho _{kli}\delta \lambda _i={\displaystyle \underset{ij}{}}\rho _{kli}(๐ด^1)_{ij}\delta \rho _j`$
Inserting this into (D1) gives the determinant form of the critical point condition
$$\underset{i}{}\frac{Y}{\rho _i}\delta \rho _i+\underset{ij}{}\underset{kl}{}\frac{Y}{\rho _{kl}}\rho _{kli}(๐ด^1)_{ij}\delta \rho _j=0.$$
(D3)
Of course, if in an application of the moment free energy method one succeeds in calculating $`Y`$ explicitly as a function of the $`\rho _i`$ alone (rather than as a function of the $`\rho _i`$ and the $`\rho _{ij}`$), then one can determine critical points simply from the condition
$`\delta Y={\displaystyle \underset{i}{}}{\displaystyle \frac{dY}{d\rho _i}}\delta \rho _i=0.`$
We have written a total derivative sign here to indicate that the dependence of $`Y`$ on the $`\rho _{ij}`$ is accounted for implicitly. |
warning/0003/gr-qc0003057.html | ar5iv | text | # Comparative Studies of Lensing Methods
## I Introduction
In this paper, three different approaches to gravitational lensing are compared in test cases with one or two lenses. The three approaches include a recently introduced exact approach , which we will take as representing correct values in our comparisons, the thin lens approximation commonly used in lensing, and a new iterative technique whose zeroth iterate is given by the thin lens approximation . An outline of these three methods will be given in Section II.
A recent paper by Frittelli, Kling, and Newman discusses lensing in a Schwarzschild geometry and compares the exact approach with the standard thin lens approximation, a second order thin lens approximation, and a strong-field version of the thin lens approximation introduced by Virbhadra and Ellis . In Schwarzschild spacetimes, it was found that the first and second order thin lens approximations fail dramatically when the light rays encounter strong gravitational fields. However, the strong-field version of Virbhadra and Ellis, which is a hybrid lensing approach using some exact and some thin lens ideas, performed remarkably well in predicting the observation angle of a given source for a given observer and lens, even when the light ray circles the lens many times. In addition, it was found that the errors in the time delays predicted by the thin lens approximation compared with the values predicted by the exact method were small for cases resembling current observational scenarios, but, as the impact parameter was reduced, the error introduced by using the thin lens approximation became appreciable.
Here we are interested in studying the accuracy of the thin lens and iterative approaches to gravitational lensing in more detail. Specifically, we wish to address two questions:
1. Are the intrinsic errors introduced by using the thin lens approximation of comparable size to the observational errors present today or in the near future?
2. Are the corrections for lens structure, given in terms of higher multipole moments or some mass distribution function, utilized by the thin lens method, comparable to the inherent errors in the thin lens approximation?
With each of these questions, we are also interested in seeing if the iterative approach provides a significant improvement in accuracy. We will not examine the accuracy of other approximate techniques in the literature .
At this point, we do not wish to imply that the work presented here should be taken as a serious attempt at modeling real lens systems. Our intention is to study the accuracy of the thin lens and iterative methods in simple test cases and to determine whether more detailed studies should be undertaken.
To address these issues, several test cases involving one or two lenses will be examined. Three different lens configurations will be compared:
a single spherically symmetric lens,
two identical lenses spatially collinear with the observer in two different lens planes,
two identical lenses located in a single lens plane.
For each of these scenarios, we will consider several comparisons:
the source location predicted by each method given the same lens configuration, observer, and observation angle,
the time delays between two rays predicted by each method given the same lens configuration, observer, and (two) observation angles.
the magnification (relative to an unlensed ray) predicted by each method given the same lens configuration, observer, and observation angle.
Section III will be devoted to these comparisons. The issue of the accuracy of the thin lens approximation was raised in Chapter 9 of , where a footnote reference is given to the work of P. Haines. One motivation of this paper is to extend the studies cited there.
In general, we find that, in the observational regime, the thin lens and iterative approaches are very accurate when applied to lensing by a single monopole lens and combination of two monopole lenses. In most cases studied, the iterative method provided only a minimal improvement over the thin lens method. However, when two lenses in separate, but closely spaced, lens planes are modeled in the thin lens approximation by a mass distribution compressed into one lens plane, observationally significant errors were found.
## II Lensing Approaches
In this section, we give brief outlines of the thin lens, exact and iterative approaches to lensing. For convenience, we will set $`G=c=1`$ in our equations, although, in our final comparisons we will return to physical units (meters, days). Our convention for the signature of the metric is $`(+,,,)`$.
### A The thin lens approximation
The thin lens approximation is the standard approach to gravitational lensing used by astrophysicists. In this subsection, we will only outline its basic premises and indicate several particular assumptions which we will use in our comparisons in Section III. Very thorough and pedagogical presentations of the thin lens methodology and its applications can be found in the excellent book by Falco, Schneider and Ehlers and a paper by Blandford and Narayan .
In the standard approach, the lens is treated as a weak perturbation of a background spacetime. For convenience, two kinds of spatial planes are introduced: a source plane, where potential sources lie, and lens planes containing lensing bodies.
Kinematically possible paths from an observer at the point O to a source at a point S in the source plane will be connected, piecewise smooth segments of geodesics in the background space. For example, if there is one lens plane, the trajectories from S to O will pass through the lens plane at the image point I, as shown in Fig. 1. The paths from S to I and from I to O are geodesics in the background metric with no influence from the lens. The only influence of the lens on the trajectory occurs at I, where the direction of the geodesic is instantaneously changed by an amount determined by a bending angle. The thin lens bending angle is a function of a mass distribution in the lens plane and the point I.
For a single, spherically symmetric lens with mass $`m`$, the bending angle, $`\alpha `$, is given by
$$\alpha =\frac{4m}{|\stackrel{}{\xi }_o|},$$
(1)
where $`|\stackrel{}{\xi }_o|`$ is the magnitude of the two dimensional vector in the lens plane locating the point I relative to the lens. This bending angle is the first order bending angle obtained in a Schwarzschild spacetime between the future and past asymptotes of a null ray connecting points at future and past null infinity. If there are many monopole lenses in the plane located at $`\stackrel{}{\xi }_i`$ with masses $`m_i`$, the bending angle is given by adding the first order contributions of the individual bending angles
$$\stackrel{}{\alpha }=\underset{i=1}{\overset{n}{}}\frac{4m_i}{|\stackrel{}{\xi }_i\stackrel{}{\xi }_o|^2}(\stackrel{}{\xi }_i\stackrel{}{\xi }_o).$$
(2)
The vectorial bending angle, $`\stackrel{}{\alpha }`$, gives the two dimensional bending angle in the lens plane. Note that if there is more that one lens plane, the bending angle in each plane will be influenced only by the lenses in that plane. Continuous mass distributions are obtained by replacing the summation in Eq. (2) by an integral and integrating a mass distribution over the entire lens plane.
If the mass distribution is known, the entire thin lens trajectory can be codified into a lens equation. If $`\stackrel{}{\theta }`$ is an angle locating the image of a lensed source and $`\stackrel{}{\beta }`$ is the โobservation angle in the absence of the lensโ given in Fig. 1, the lens equation for one lens plane is
$$\stackrel{}{\beta }=\stackrel{}{\theta }\frac{D_{ls}}{D_s}\stackrel{}{\alpha }.$$
(3)
An important quantity in lensing is the time which elapses between the emission of the light ray and its interception by the observer, or the time of arrival. Although we will not use the ideas here, the time of arrival serves as a Fermat potential from which one can derive the lens equation, Eq. (3). Individual arrival times are not an observable, but the time delay between two images of the same source is an important quantity which has been measured in several lens systems . (The time delay can be used, with other observations, to determine the Hubble constant.)
Returning to the general case with many lenses in many lens planes, the time of arrival in the thin lens approximation is given by
$$t=_{SIO}(1+2U(l))๐l,$$
(4)
where $`U(l)`$ is the Newtonian potential of the mass distribution, and the integral is taken along the thin lens trajectory parameterized by an Euclidean length $`l`$ from the source at $`S`$ to through the image point at $`I`$ to the observer at $`O`$. For a collection of masses, the Newtonian potential is
$$U(l)=\underset{i=1}{\overset{n}{}}\frac{m_i}{|\stackrel{}{r}(l)\stackrel{}{r}_i|}.$$
(5)
In Eq. (5), $`\stackrel{}{r}_i`$ is a three dimensional vector locating lenses relative to some origin while $`\stackrel{}{r}(l)`$ is the vector locating points along the thin lens path parameterized by a Euclidean length along the trajectory. Time delays are given by the difference in two such times.
In this paper, we will choose Minkowski spacetime as our background. We do not anticipate that there would be significant differences in our results using Robertson-Walker or on-average Robertson-Walker metrics as the background spacetime, although we have not examined this question.
### B Exact lensing
The key difference between the exact approach to gravitational lensing and the thin lens approximation is that, in the exact approach, the lens is fully incorporated into a metric satisfying Einsteinโs equations. In this way, no background / lens splitting is introduced and no quantities defined in the thin lens method as โin the absence of a lensโ have meaning.
Lensing information in the exact approach is obtained by integrating the null geodesic equations of the metric . A particular parametric form of the null geodesic equations are defined to be the lens and time of arrival equations. It can be shown that in a Schwarzschild spacetime, the lens and time of arrival equations can be expressed parametrically in local, timelike coordinates as
$`t`$ $`=`$ $`T(R,\theta _o,\varphi _o,x_o^a)`$ (6)
$`x^i`$ $`=`$ $`X^i(R,\theta _o,\varphi _o,x_o^a)`$ (7)
where $`(t,x^i)`$ label points on the past light cone of an observer located at $`x_o^a`$ . The two โangular parameters,โ $`(\theta _o,\varphi _o)`$, represent the direction on an observerโs celestial sphere where the image is observed, and $`R`$ can be taken as a physical distance to the source such as the angular-diameter distance. Equation (6) is defined to be the exact time of arrival equation while Eqs. (7) are the exact lens equations.
The first comparison which we will consider in this paper is lensing in Schwarzschild spacetimes. The time of arrival and lens equations for a Schwarzschild spacetime can be found in closed form by integrating the null geodesics of the Schwarzschild metric using its symmetries ; however, in the current work, we will not use these closed form expressions for the lens and time of arrival equations. Instead, we will solve the null geodesic equations by forming a Hamiltonian,
$$H(x^c,p_c)=g^{ab}(x^c)p_ap_b,$$
(8)
and numerically solving Hamiltonโs equations of motion. Null geodesics are obtained when the initial conditions, $`(x_o^a,p_a^o)`$, satisfy
$$H(x_o^c,p_c^o)=0.$$
(9)
We will refer to the numerical integrations of the geodesic equations as the โexactโ approach or method.
By performing a coordinate transformation,
$`t`$ $`=`$ $`t`$ (10)
$`r`$ $`=`$ $`\sqrt{x^2+y^2+z^2}`$ (11)
$`\theta `$ $`=`$ $`\mathrm{arctan}{\displaystyle \frac{\sqrt{x^2+y^2}}{z}}`$ (12)
$`\varphi `$ $`=`$ $`\mathrm{arctan}{\displaystyle \frac{y}{x}},`$ (13)
we can write the Schwarzschild metric as
$$g^{ab}(x^a)=\left(\begin{array}{cccc}1/(1\frac{2m}{r})& 0& 0& 0\\ 0& 1+\frac{2mx^2}{r^3}& \frac{2mxy}{r^3}& \frac{2mxz}{r^3}\\ 0& \frac{2mxy}{r^3}& 1+\frac{2my^2}{r^3}& \frac{2myz}{r^3}\\ 0& \frac{2mxz}{r^3}& \frac{2myz}{r^3}& 1+\frac{2mz^2}{r^3}\end{array}\right).$$
(14)
These coordinates are useful for comparing the โexactโ method to the thin lens and iterative methods. Hence, we solve Hamiltonโs equations of motion in a $`(t,x,y,z)`$ coordinate system which is adapted to comparisons with the thin lens and iterative approaches.
We are also interested in comparing the three lensing methodologies in situations with more than one lens. Since there are no exact solutions to Einsteinโs equations to meet these lensing configurations, one can not apply the exact methodology. For these cases, we will consider approximate metrics whose null geodesics are solved using Hamiltonโs equations of motion. In these cases, we will call the numerical solution to Hamiltonโs equations for null geodesics of the approximate metric the โexactโ time of arrival and lens equations.
### C Iterative approach
The iterative approach seeks to improve upon the thin lens approximation and can be applied to any approximate solution to Einsteinโs equations which is close to a spacetime in which the exact method can be employed. In this paper, we will focus on the first iterate only, although higher iterates can be obtained. Details on the iterative approach, including equations for the first iterate method applied to a Schwarzschild spacetime, can be found in .
The general method is to assume that the spacetime of interest, $`(M,g^{ab})`$, is close to some spacetime, $`(M,g_o^{ab})`$, where the geodesic equations can be solved exactly. This means that we can write the metric $`g^{ab}`$ as
$`g^{ab}(x^c)=g_o^{ab}(x^c)+h^{ab}(x^c)`$
where the components of $`h^{ab}(x^c)`$ are small.
One begins the iterative method by forming the Hamiltonians in both spacetimes,
$$H(x^c,p_c)=g^{ab}(x^c)p_ap_b=g_o^{ab}(x^c)p_ap_b+h^{ab}(x^c)p_ap_b$$
(15)
$$H_o(x^c,p_c)=g_o^{ab}(x^c)p_ap_b,$$
(16)
and solving the Hamilton-Jacobi equation in $`(M,g_o^{ab})`$:
$$g_o^{ab}(x^c)\frac{F}{x^a}\frac{F}{x^b}+\frac{F}{\lambda }=0.$$
(17)
In spacetimes in which the geodesic equations can be solved, one can always find a solution to the Hamilton-Jacobi equation of the form
$$F=F_o(x^c,P_c,\lambda )$$
(18)
in which $`\lambda `$ is a parameter and $`P_c`$ are four constants. This function can be taken as the generating function for a parameter dependent, canonical transformation,
$$(x^c,p_c)(X^c,P_c).$$
(19)
If $`(M,g_o^{ab})`$ is taken to be Minkowski spacetime, $`(M,\eta ^{ab})`$, the solution to the Hamilton-Jacobi equation, Eq. (17) is
$$F_o(x^c,P_c,\lambda )=x^aP_a\eta ^{ab}P_aP_b\lambda ,$$
(20)
and the canonical transformation to the coordinates $`(X^c,P_c)`$ is given by
$`x^a`$ $`=`$ $`X^a2\eta ^{ab}P_b\lambda `$ (21)
$`p_a`$ $`=`$ $`P_a.`$ (22)
When this canonical transformation is applied to the Hamiltonian in $`(M,g^{ab})`$, the transformed Hamiltonian takes a particularly simple form
$$H(x^c,p_c)H^{}(X^cP_c,\lambda )=h^{ab}(X^c,P_c,\lambda )P_aP_b.$$
(23)
Hamiltonโs equations for geodesics in the $`(X^a,P_a)`$ coordinates are
$`\dot{X}^a`$ $`=`$ $`2h^{ab}(X^c,P_c,\lambda )P_b+\left({\displaystyle \frac{h^{bc}}{P_a}}(X^c,P_c,\lambda )\right)P_bP_c\mathrm{\Xi }^a(X^a,P_a,\lambda )`$ (24)
$`\dot{P}_a`$ $`=`$ $`\left({\displaystyle \frac{h^{bc}}{X^a}}(X^c,P_c,\lambda )\right)P_bP_c\mathrm{\Pi }_a(X^a,P_a,\lambda ).`$ (25)
No approximations have been made to obtain these equations.
We wish to solve Hamiltonโs equations, Eqs. (25), by iteration. For the zeroth iterate, we must specify eight functions
$`X_0^a`$ $`=`$ $`X_0^a(X_o^a,P_a^o,\lambda )`$ (26)
$`P_a^0`$ $`=`$ $`P_a^0(X_o^a,P_a^o,\lambda ),`$ (27)
and substitute these functions of $`\lambda `$ and initial conditions, $`x_o^a=X_o^a`$ and $`p_a^o=P_a^o`$, into the right hand side of Hamiltonโs equations, Eqs. (25). (Care should be taken to choose the eight functions serving as the zeroth iterate close to the true description of the path of the null geodesic.) The first iterate is obtained by direct integration on $`\lambda `$:
$`X_1^a(X_o^a,P_a^o,\lambda )`$ $`=`$ $`X_o^a+{\displaystyle _0^\lambda }๐\lambda ^{}\mathrm{\Xi }^a(X_0^a,P_a^0,\lambda ^{})`$ (28)
$`P_a^1(X_o^a,P_a^o,\lambda )`$ $`=`$ $`P_a^o+{\displaystyle _0^\lambda }๐\lambda ^{}\mathrm{\Pi }_a(X_0^a,P_a^0,\lambda ^{}).`$ (29)
Likewise, the $`n`$th iterate is obtained by placing $`(X_{n1}^a(X_o^a,P_a^o,\lambda ),P_a^{n1}(X_o^a,P_a^o,\lambda ))`$ into the right hand side of Eqs. (25) and integrating up.
The $`n`$th iterate solution to Hamiltonโs equations in the original spacetime coordinates, $`(x^a,p_a)`$, is obtained by placing $`(X_n^a(X_o^a,P_a^o,\lambda ),P_a^n(X_o^a,P_a^o,\lambda ))`$ into the canonical transformation, Eq. (21) and (22):
$`x_n^a(X_o^a,P_a^o,\lambda )`$ $`=`$ $`X_n^a(X_o^a,P_a^o,\lambda )2\eta ^{ab}P_b^n(X_o^a,P_a^o,\lambda )\lambda `$ (30)
$`p_a(X_o^a,P_a^o,\lambda )`$ $`=`$ $`P_a^n(X_o^a,P_a^o,\lambda ).`$ (31)
In these equations, $`(X_o^a=x_o^a,P_a^o=p_a^o)`$ are the initial values for the approximate geodesic. When these initial conditions satisfy the null condition on the Hamiltonian in Eq. (15),
$$H(X_o^a=x_o^a,P_a^o=p_a^o,\lambda =0)=g^{00}(x_o^a)(p_0^o)^2+g^{ij}(x_o^a)p_i^op_j^o=0,$$
(32)
the geodesics are approximately null. Solving Eq. (32) for $`p_0^o`$ yields
$$p_0^o=\sqrt{\frac{g^{ij}(x_o^i)p_i^op_j^o}{g^{00}(x_o^a)}}.$$
(33)
We will make use of this expression below.
To make a deeper connection with lensing, we will take the thin lens path as the zeroth iterate. As an example, we show the explicit zeroth iterate for the case of one spherical lens in one lens plane.
For one spherically symmetric lens, the thin lens path would be given by
$`x_{(1)}^i`$ $`=`$ $`x_o^i2\delta ^{ij}p_j^o\lambda 0<\lambda <\lambda _1`$ (34)
$`x_{(2)}^i`$ $`=`$ $`x_{(1max)}^i2\delta ^{ij}\stackrel{~}{p}_j^o\lambda 0<\lambda ,`$ (35)
where $`x_{(1)}^i`$ describes the first leg (to the lens plane), $`x_{(2)}^i`$ describes the second leg away from the lens plane and $`(x_o^i,p_i^o)`$ are constant initial conditions. If we locate the observer at $`|z_o|`$ on the $`\widehat{z}`$ axis, we can use spherical symmetry to consider geodesics in the $`\widehat{x}`$-$`\widehat{z}`$ plane. Then the $`z=0`$ plane is the lens plane, and the value of $`\lambda _1`$ is
$$\lambda _1=\frac{|z_o|}{2p_z^o}.$$
(36)
We also have that
$$x_{(1max)}^i=x_o^i\frac{\delta ^{ij}p_j^o|z_o|}{p_z^o}.$$
(37)
Using the bending angle, $`\alpha =2m/|x_{(1max)}^i|`$, the $`\stackrel{~}{p}_j^o`$ will be determined up to scaling through the Minkowski spatial inner product between $`p_j^o`$ and $`\stackrel{~}{p}_j^o`$:
$$\mathrm{cos}\alpha =\frac{(p_x^o)(\stackrel{~}{p}_x^o)+(p_z^o)(\stackrel{~}{p}_z^o)}{|p_i^o||\stackrel{~}{p}_i^o|}.$$
(38)
If we multiply and divide the right hand side in Eq. (38) by $`1/(p_z^o\stackrel{~}{p}_z^o)`$ and define
$`p={\displaystyle \frac{p_x^o}{p_z^o}}\mathrm{and}\stackrel{~}{p}={\displaystyle \frac{\stackrel{~}{p}_x^o}{\stackrel{~}{p}_z^o}},`$
Eq. (38) is equivalent to a quadratic equation for $`\stackrel{~}{p}`$ in terms of $`p`$ and $`\alpha `$. Solving this equation gives
$$\stackrel{~}{p}=\frac{p\pm \mathrm{cos}\alpha \mathrm{sin}\alpha \sqrt{(1+p^2)^2}}{\mathrm{cos}^2\alpha p^2\mathrm{sin}^2\alpha }.$$
(39)
We now return to the issue of the value of $`p_0^o=P_0^o`$. In this paper, the iterative method will only be applied to stationary spacetimes. Hence, the timelike coordinate, $`t`$, will not appear in the original Hamiltonian, Eq. (15), and $`T`$, the canonically transformed variable, will not appear in the transformed Hamiltonian, Eq. (23). As $`T`$ is cyclic, the time equation, $`(\dot{T})`$, will separate from the spatial equations. Moreover, because $`T`$ does not appear in the Hamiltonian,
$`\dot{P}_0={\displaystyle \frac{H^{}}{T}}`$
will be identically zero so that $`P_0=p_0`$ is constant. The value of this constant must be chosen to make the trajectory null, as in Eq. (33). However, there is an inherent scaling freedom of the null vector which permits us to define new four-momenta, $`p_a^{}_{}{}^{}o`$ which are a constant multiple of the old one.
It is customary to fix the freedom in the scaling of the null vector by defining $`p_0^o^{}=1`$. Then in the case of a single monopole lens, the equation
$$p_0^o1=\sqrt{\frac{g^{ij}(x_o^a)p_i^op_i^o}{g^{00}(x_o^a)}},$$
(40)
gives an equation for the initial $`p_z^o`$ in terms of the $`p_x^o`$ and the initial point.
As we take the thin lens trajectory as the zeroth iterate, we will choose to set the value of $`p_0^o`$ to one at the initial point (the observer, $`O`$) and again in each lens plane. In this way, $`p_0^o=\stackrel{~}{p}_0^o=1`$, and the relative scaling of $`(p_x^o,p_z^o)`$ and $`(\stackrel{~}{p}_x^o,\stackrel{~}{p}_z^o)`$ is uniquely determined using Eq. (40).
Summarizing, the spatial part of the zeroth iterate for a one lens system is given by the thin lens path, Eq. (35). The observer is located at the initial point $`x_o^a`$. For a given value of $`p_x^o`$, $`p_z^o`$ is uniquely determined by the condition $`p_0^o=1`$, from Eq. (40). The new $`(\stackrel{~}{p}_x^o,\stackrel{~}{p}_z^o)`$ are determined using the bending angle as in Eq. (39) and the condition $`\stackrel{~}{p}_0^o=1`$.
So far, our discussion has only considered cases with axial symmetry and one lens plane. If there is more than one lens plane, the procedure we have described above is extended to each lens plane. If there is no axial symmetry, there will be a complementary equation to the bending angle relation, Eq. (38), which can be used to fix the three spatial components of the momenta in an analogous way to what we have presented here.
In the cases we will study, the time coordinate is cyclic and the time equation in the original phase space variables is simply the integral equation
$$t=t_o+2p_0^o(1+2U(x^a(\lambda )))๐\lambda ,$$
(41)
where $`U(r)`$ is the Newtonian potential of the mass distribution. The integral is taken over the path as a function of the parameter $`\lambda `$. The first iterate time is obtained when the path inserted into Eq. (41) is the zeroth order, thin lens path, $`x^a(\lambda )=x_0^a(\lambda )`$. Hence the first iterate time is precisely the value obtained in the standard thin lens approximation. However, as we will find the first iterate trajectories, we can also find the second iterate time values by evaluating the integrand in Eq. (41) along the first iterate, $`x^a(\lambda )=x_1^a(\lambda )`$.
A conceptual problem arises in computing the iterative time of arrival if one uses the formula
$$t_n=t_o+2p_0๐\lambda (\mathrm{\hspace{0.17em}1}+U(x_{n1}^a(\lambda )))$$
(42)
for the $`n`$th iterate time. In general, the parameter $`\lambda `$ should be an affine parameter along a null geodesic. We note that if Hamiltonโs equations are solved exactly, fixing $`p_0^o`$ as in Eq. (33) ensures that the value of the Hamiltonian will be zero at all points along the path and $`\lambda `$ will be an affine parameter. However, in the iterative method, the geodesic equations are not solved exactly, and the value of the Hamiltonian will slowly drift away from zero. Hence, as $`\lambda `$ grows, it fails to be an affine parameter along a null geodesic.
A way to force the Hamiltonian to be zero, and hence $`\lambda `$ to be null affine parameter, is to allow $`p_0`$ to be a function along the trajectory given by
$$p_0=p_0^n(\lambda )=\sqrt{\frac{g^{ij}(x_n^a)p_i^np_j^n}{g^{00}(x_n^a)}},$$
(43)
where $`(x_n^a,p_a^n)`$ are the $`n`$th iterate values. This proposal leads to a conflict between maintaining the null value of the Hamiltonian along the $`n`$th iterate trajectory
$$H(x_n^a,p_a^n,\lambda )=0$$
(44)
and Hamiltonโs equation,
$$\dot{p}_0=0.$$
(45)
Since we are dealing with static cases, a consistent proposed solution is to take $`p_0`$ as constant in the spatial Hamiltonโs equations, but allow $`p_0`$ to vary as in Eq. (43) when computing times. This solution disentangles the two competing problems given in Eq. (44) and Eq. (45) by obeying Hamiltonโs equations when integrating the spatial part of the geodesic and also obeying the null condition in computing the times (which is very sensitive to integrating over an affine parameter).
Hence, when comparing the iterative time delays to the โexactโ and thin lens delays, we will consider the time of arrivals given by
$$t_2=t_o+\mathrm{\hspace{0.17em}2}p_0^1(\lambda )(1+2U(r_1))๐\lambda .$$
(46)
We will show that this formula gives very accurate predictions for time delays in our comparisons.
## III Comparison of Lensing Approaches
In this section we present the results obtained in the comparison of time delays, source locations and image magnifications predicted by the thin lens approximation and the iterative method for several different lens models. The comparisons are made with respect to the numerical integration of the exact geodesic equations of the spacetime metric defined by the given model. For the iterative method, we will be interested in the first iterate only.
There are four subsections in this section. In the first, we give details regarding the comparisons we will be discussing. We then group our comparisons in three sets. First, we consider lensing by a single spherical lens, or the Schwarzschild geometry. Next, we consider multiple lensing by single monopole lenses collinear with the observer in different lens planes. Finally, we consider lensing by two lenses in the same lens plane. In our plots, all angles are given in arc seconds and all times are given in days.
### A Notes about comparisons
We will be comparing the predictions of the thin lens, iterative and โexactโ approaches to lensing for time delays, source locations and image magnification. In this subsection, we give some details about how these comparisons are performed. We will refer to the spatial axis connecting the lens and observer as the optical axis.
First, we note that the exact solution to the geodesic equation produces an infinite number of images , but that the thin lens approximation predicts only two of these images for the case of a single lens (referred to as primary images). This feature is shared by the first iterate method, since it corresponds to the next step in the perturbative series whose zeroth order is given by thin lens approximation (the first iterate method should give sensible predictions as long as its trajectory remains close to the thin lens trajectory). However, it is not difficult to choose the two primary exact images corresponding to the thin lens and first iterate images because these primary images are widely separated (in angular location) from the secondary images which circle the lens one time.
Throughout our comparisons, we will refer to an observation angle, $`\theta `$. This angle is computed by taking the inner product between the spatial part of the initial momentum vector, $`p_i^o`$, at the observer and the spatial vector pointing to the lens from the observerโs location, $`x_o^a`$. Formally, this must be done using the spatial metric describing the model. However, as we will always be taking this inner product at a large distance from the lens, it is appropriate to take the observation angle, $`\theta `$, as the ratio of the $`\widehat{x}`$ and $`\widehat{z}`$ components of the initial momentum:
$`\theta ={\displaystyle \frac{p_x^o}{p_z^o}}.`$
To compare the time delays predicted by the three methods, we must compute the two trajectories from each method, integrate the arrival time function along each trajectory, and subtract the two values we obtain. We begin by choosing two initial angles, $`(\theta _1,\theta _2)`$, one on each side of the lens. These angles are chosen such that trajectories with these initial conditions intersect at a reasonable distance beyond the last lens; usually, this distance is chosen as approximately the same as the distance between the observer and the first lens. For each path, we compute the time of arrival and subtract the two times to get a time delay. We will then hold one angle, $`\theta _2`$, fixed while varying $`\theta _1`$. This allows us to consider the time delay as a function of $`\theta _1`$ for a fixed $`\theta _2`$.
In practice, finding the time delays is a very difficult calculation, as the arrival times for each trajectory will agree in roughly their first 12 digits (at our scales). The comparison between the methods is even more difficult, as the time delays from the three methods tend to agree to about four digits. Hence, to resolve a difference between the thin lens, iterative, and โexactโ predictions for the time delays, we must know the arrival time to approximately 16 digits of accuracy.
To compute the source location, $`\beta `$, for a given observation angle, $`\theta `$, we choose a value of $`\theta `$ and a final distance along the optical axis from the observer, $`D_s`$. If the optical axis is the $`\widehat{z}`$ axis and the observer is located at $`z=|z_0|`$, we then place a plane at $`z=D_s|z_0|`$ which will be the โsource plane.โ We then compute, for a given initial condition, $`\theta `$, the interception point in the source plane, $`\stackrel{}{\eta }`$. The value of $`\beta `$ is defined to be
$$\beta =\frac{|\stackrel{}{\eta }|}{D_s}.$$
(47)
We will use this definition for the source location in all three models. As $`\theta `$ is varied, we will obtain $`\beta (\theta )`$.
Magnifications are defined as $`\frac{\theta }{\beta }`$, or the inverse slope of the $`\beta `$ versus $`\theta `$ graph. We will compute the magnifications from the data we obtain in our $`\beta `$-$`\theta `$ comparisons. Note that this magnification is not directly observable; we discuss it only because it plays a role in the literature. With some additional work, we could compute, for two given images, $`(\theta _1,\theta _2)`$, the relative magnification,
$`\mu _{12}={\displaystyle \frac{\left(\frac{\theta }{\beta }\right)_{\theta _2}}{\left(\frac{\theta }{\beta }\right)_{\theta _1}}},`$
which is observable. We plan to return to this possible comparison in future work.
In the two lens models, the thin lens approximation predicts four images. Therefore, in the calculation of the time delays we can distinguish three qualitatively different situations. As it is illustrated in Fig. 2, there is a range for $`\theta _1`$ and $`\theta _2`$ for which the two rays do not cross the optical axis between the two lenses before converging at the observerโs position (range A), a range for which only one of the rays crosses the axis between the lenses (range B), and, finally, a range in which both of the rays cross the axis before they meet at the observerโs position (range C). When looking at magnifications and image positions, we can compare the three methods in two cases: 1) rays which do not cross between the two lenses and 2) rays which do cross the optical axis between the lenses.
### B One spherical lens
Here we discuss the comparisons between the first iterate, thin lens and โexactโ predictions for various observables when there is one spherically symmetric lens. For the โexactโ predictions, we consider the numerical integration of the geodesic equations of the exact Schwarzschild metric, as specified in Section II.B. As mentioned, Minkowski spacetime will be considered the background spacetime for the iterative and thin lens approaches.
For our comparisons, we will take a lens $`4`$ billion light years (ly) away from the observer with a mass of approximately $`2.5\times 10^{12}M_{}`$. While we are not concerned with cosmological models here, this distance scale is reasonable for current lensing studies.
In Fig. 3a we show the position of the source, $`\beta `$, as a function of the image position, $`\theta `$, (see Fig. 1) calculated using the exact numerical integration of the geodesics equations. The range in $`\theta `$ has been chosen in agreement with observed image angles in systems with similar characteristics as the one represented by our model, and the value of $`D_s`$ has been set equal to the observer-lens spacing. The absolute error in $`\beta `$ between the exact numerical integration and the thin lens approximation and the first iterate method are shown in Fig. 3b. The discrepancies between the two methods are of about $`10^5`$ arc sec. In Fig 4a, we show the โexactโ magnification $`\mu =\frac{\theta }{\beta }`$ as a function of the image position $`\theta `$. The relative error,
$`\mathrm{\Delta }\mu _{tl,it}={\displaystyle \frac{\mu _{ex}\mu _{tl,it}}{\mu _{ex}}},`$
in the predicted magnification by the two approximate methods is shown in Fig 4b. The errors here are very small.
In the case of a single spherically symmetric lens, we were not able to resolve the difference in the time delay error between the thin lens and first iterate trajectories. Our calculations showed that this error was indeed quite small, and that for a lens with mass $`2.5\times 10^{12}M_{}`$ at $`4`$ billion ly from the observer, the error in the thin lens and iterative methods was less than $`0.2`$ days when the โexactโ time delay was $`400`$ days.
### C Two lenses in different lens planes
In this subsection, we will consider lensing by two identical lenses. We choose to study two different cases: when the distance between the two lenses is on the same order of magnitude as the distance between the observer and the first lens and when the distance between the two lenses is small compared to the distance between the observer and the first lens. We will examine the case where the lenses are far apart first.
$`1)`$ Large separation
In our first comparisons, we consider an observer $`3`$ billion ly away from the first lens and set the distance between the two lenses equal to the distance between the observer and the first lens. The mass of the two lenses is approximately $`1.9\times 10^{12}M_{}`$. Figure 5a shows the โexactโ image location, $`\beta `$, as a function of observation angle, $`\theta `$, when the light ray does not cross between two lenses and $`D_s=9`$ billion ly, or three times the spacing between the first lens and observer. As before, the error in the thin lens and first iterate methods, shown in Fig. 5b, are small. At a fairly large observation angle, $`4.3^{\prime \prime }`$ from the optical axis, the error in the thin lens method is about $`3\times 10^5`$ arc sec. The โexactโ magnification and errors in the thin lens and iterative methods for this lens configuration are shown in Fig. 6.
For the same lensing configuration, the โexactโ source location and the error in the thin lens and iterative methods when the light ray crosses the optical axis is shown in Fig. 7. In this case, the light ray passes much closer to the lens, and we see that the first iterate is slightly better than the thin lens in predicting the source location. The corresponding magnifications are plotted in Fig. 8.
With the lens separation equal to the distance between the observer and the first lens, the time delays predicted by the thin lens and iterative method are very accurate. As in the one lens case, we were unable to resolve a difference in the time delays due to the high precision required in any of the three possible ray combinations from Fig. 2. Our calculations show that the error in the thin lens and first iterate methods was less than $`0.1`$ days for a time delay around $`500`$ days.
$`2)`$ Small separation
Similar results were obtained when the distance between the two lenses was small. As an example, we will examine the case where the observer and two lenses lie along the same optical axis, the mass of each lens is $`2.5\times 10^{12}M_{}`$, the distance to the first lens from the observer is $`4`$ billion ly and the distance between the two lenses is $`4`$ million ly. This second distance is about twice the distance between our galaxy and Andromeda, so that our lensing configuration represents a pair of lenses at roughly the same distance from the observer. Thus, we may think of this example as corresponding to direct lensing by two members of the same group. In this case, it makes sense to choose $`D_s`$ as twice the distance to the first lens, $`D_s=8`$ billion ly.
When the lenses are so close together, it does not make sense to consider rays crossing between the two lenses; to pass between the lenses, the observation angle must be less than $`0.125^{\prime \prime }`$. Because this observation angle does not look reasonable, we will not compare the thin lens, iterative and exact methods in this paper in this range of $`\theta `$, although we note that in an analogous case of microlensing such comparisons may be important.
As in our previous comparisons, we show the โexactโ source angle as a function of the observation angle and the error in the thin lens and first iterate approaches in Fig. 9 for the case where the two lenses are close together. We note that the error in the thin lens method is approximately $`\beta _{ex}\beta _{tl}5.5\times 10^5`$ arc sec, while the error in the iterative method is approximately $`\beta _{ex}\beta _{it}5.0\times 10^5`$ arc sec. These errors are somewhat larger than the errors when the lens planes are widely separated.
A similar result is found in the magnifications, shown in Fig. 10. Here, we note that the inaccuracy in both methods is about twice the inaccuracy in the case where the lenses are widely separated.
We did not detect an observable error in the time delays for the thin lens or iterative methods for this scenario. For a โexactโ time delay of $`400`$ days, the thin lens and iterative methods were accurate to less than $`0.1`$ days for observation angles around $`2.475^{\prime \prime }`$.
Because the lenses are so close together, one may think that it is appropriate to treat them as a single lens located in the middle of the two with a total mass equal to the sum of the values of the two masses. We will refer to this model as the โ2 lens, single lens plane model.โ When the thin lens approximation is applied to the โ2 lens, single lens plane model,โ the error in the time delay is significant at this separation. The curve in Fig. 11b represents the error in the time delay predicted by the thin lens approximation when the lensing configuration is treated as one lens with the total mass located directly between the two lenses. Here, one observation angle was fixed at $`2.475^{\prime \prime }`$ and the second angle varies as shown. The exact time delay is plotted in Fig. 11a.
The errors present in the โ2 lens, single lens plane modelโ are significant and are of the same order of magnitude as current observational abilities. Hence, it appears that lens structure extending along the optical axis connecting the observer and first lens can adversely affect the accuracy of the thin lens methodology at todayโs observational level when the structure is collapsed into a single lens plane. We will discuss this issue further in Sec. IV.
### D Two lenses in the same lens plane
As a final comparison, we consider two lenses in the same lens plane. Here we will take each lens to have a mass of $`1.25\times 10^{12}M_{}`$ and will set the lens plane $`4`$ billion ly away from the observer. This case resembles the one lens system in that the distances are the same but the mass has been split. We choose a separation of $`4000`$ ly between the lenses. For our comparisons of $`\beta `$ and magnification, we will take $`D_s`$ to be twice the spacing between the lens and observer.
Figure 12 shows the โexactโ plot of source angle, $`\beta `$, versus observation angle, $`\theta `$, and the error in $`\beta `$ for the thin lens and iterative method. The โexactโ magnifications and relative errors are shown in Fig. 13. As in the case of the two lenses close together, there is a slight difference in the accuracy of the thin lens and iterative methods. Again, no measurable error was found in the time delays predicted by the thin lens and iterative methods.
## IV Discussion
We have performed a careful examination of the accuracy of two lensing approximations, the thin lens and iterative methods, in scenarios with one and two monopole lenses. In general, we find that source locations, time delays and magnifications computed using the first iterate are more accurate than those computed with the thin lens.
Both methods are accurate beyond the current level of observational error when the deflector was a single spherical lens or two lenses in all cases tested. The cases we studied in this paper tended to involve rather massive lenses. Since the thin lens and iterative methods generally become more accurate as the mass is decreased holding the distances the same, we feel that it is likely that lensing by objects with smaller masses than those considered here will be well described by both the thin lens and iterative methods.
It was found that when two closely separated lenses on the optical axis were modeled in the thin lens approximation by a mass distribution in one lens plane (the โ2 lens, single lens plane modelโ), observationally significant errors arise. Because these errors approach zero in the limit that the distance between the two lens planes goes to zero, the important observational question is whether there are observational scenarios where the depth of the mass distribution along the line of sight is too large to be modeled by one lens plane.
The failure of the โ2 lens, single lens plane modelโ to accurately predict observable quantities for some separations raises two interesting questions for lensing. First, are there lensing scenarios where multiple members of a group lens a source? In this case, careful observational work needs to be done to determine the relative spacing of the members of the group, for if the spacing is too large, significant errors may be introduced.
It seems that when there is a three dimensional mass distribution (a lens with structure), collapsing the distribution into a single lens plane may lead to an approximation which fails at todayโs level of observational accuracy. Further studies are needed to determine if two dimensional continuous mass distributions of the type used in lens modeling are affected at the same level as the collection of monopole lenses studied here. This issue is important because two dimensional continuous mass distributions are used to predict various cosmological parameters, and the inability to correctly model time delays, source positions and image magnifications will lead to inaccuracy in the prediction of fundamental constants from lensing.
As a second question, we are interested in how our results apply to microlensing by binary systems. It is estimated that nearly ten percent of microlensing events will be microlensing by binary systems. At various points in time, the rotating bodies will resemble either two lenses in different lens planes or two lenses in the same lens plane, which are similar to our case studies. One key difference is that, in general, the mass to distance ratio in microlensing will be smaller than the ratios we have studied. This will tend to reduce the error we detected in the thin lens method. On the other hand, as the source moves across the sky, the light rays from the source come very close to the binary lens. We found a rather large error in the thin lens method when the distance of closest approach was on the same order of magnitude as the separation between the lenses. Hence, we do not know what the accuracy of the thin lens method will be when applied to microlensing by binaries. We will study this issue in future work.
In summary, we have shown that the intrinsic errors of the thin lens approximation fail to approach todayโs level of observational error (approximately one milli arc sec for angles in the visible band and one day for time delays) by approximately two orders of magnitude for one or two monopole lenses. The inherent errors in the iterative method were consistently smaller than the errors of the thin lens method, although these errors were of the same order of magnitude in almost all cases.
On the other hand, when a single lens plane is used to model two closely separated lenses in different lens planes, significant errors did arise in time delays. This suggests that the โ2 lens, single lens plane modelโ should be applied very carefully in observational cases; one should be careful to check that lens structure does not extend a significant fraction of the distance along the line of sight between the lens and observer.
Even though the inherent errors in the thin lens approximation are not a significant fraction of the observational errors in the cases we studied, it is not inconceivable that the observational accuracy will improve over time to a point where more sophisticated approaches are required for modeling lens trajectories. The iterative method provides one such improvement over the thin lens and seems to be accurate in all the cases we have studied.
Acknowledgments
The authors would like to thank David Turnshek, Al Janis, Simonetta Frittelli and Jurgen Ehlers for their helpful advice and suggestions. Alejandro Perez would like to thank FUNDACION YPF. This work was supported under grants Phy 97-22049 and Phy 92-05109. |
warning/0003/math0003073.html | ar5iv | text | # Loop observables for ๐ตโข๐น theories in any dimension and the cohomology of knots
## 1. Introduction
Knot invariants can be obtained as expectation values of Wilson loops (i.e., traces of holonomies) in ChernโSimons theory .
The same result can be obtained in the $`3`$-dimensional $`BF`$ theory โwith a cosmological termโ .
The nice feature of $`BF`$ theory , as opposed to ChernโSimons theory, is that it can be defined in any dimension and always has a quadratic term around which one can start a perturbative expansion.
On the other hand, apart from the $`3`$-dimensional case , it is rather difficult to find nontrivial observables for $`BF`$ theories (see for the search of surface observables in the $`4`$-dimensional case).
However, the formulae for the perturbative expansion of the expectation value of a Wilson loop in the $`3`$-dimensional ChernโSimons theory (e.g., in the approach of ) allow for a natural generalization in any dimension . The invariants defined in this way are cohomology classes of Vassiliev finite type on the space of imbeddings of $`S^1`$ into $`^n`$. As in the $`3`$-dimensional case, they are moreover related to certain โgraph cohomologies,โ as originally suggested by Kontsevich .
This led us to look for loop observables for $`BF`$ theories in any dimension whose expectation values should yield the above invariants.
In this paper we introduce these observables โon shellโ (i.e., upon using the equation of motions) and describe some results about their off-shell extension (which relies on the use of the BatalinโVilkovisky formalism). For all technical details, as well as for the complete proofs of the four Theorems contained in this paper, we refer to .
A very interesting feature of the simplest of these observables (as well as of the cohomology classes described in ) is that, despite of the dimension, they are always as if we were dealing with the $`3`$-dimensional $`BF`$ theory with a cosmological term, which is tantamount to considering the ChernโSimons theory. This is our interpretation of Wittenโs ideas about a ChernโSimons theory for strings .
###### Acknowledgment.
We thank R. Longoni for carefully reading the manuscript and for many useful discussions.
## 2. $`BF`$ theories
The $`n`$-dimensional $`BF`$ theory is a topological quantum field theory defined in terms of a connection $`1`$-form $`A`$ over a principal $`G`$-bundle $`PM`$, with $`dimM=n`$, and a tensorial $`(n2)`$-form $`B`$ of the $`\mathrm{ad}`$ type.
Following the ideas of , we begin with a geometrical description of gauge invariant functionals of $`A`$ and $`B`$ taking values in the free loop space $`\mathrm{L}M`$.
The basic functional is of course the Wilson loop $`\mathrm{Tr}_\rho \mathrm{Hol}`$. Here we follow the following nonstandard
###### Convention.
$`\mathrm{Hol}(\gamma ;A)`$ denotes the group element associated to the $`A`$-parallel transport along the loop $`\gamma :[0,1]M`$ from $`\gamma (1)`$ to $`\gamma (0)`$. (The usual holonomy is the inverse of our $`\mathrm{Hol}`$.)
We then associate to each smooth loop $`\gamma `$ in $`M`$ and to any connection $`A`$ the $`A`$-horizontal lift of $`\gamma `$: $`[0,1]t\gamma _A(t)`$.
If we saturate the $`(n2)`$-form $`B`$ with the tangent vector $`\dot{\gamma }_A(t)`$, we obtain an $`(n3)`$-form $`B(\dot{\gamma }_A(t))`$ defined along the horizontal path $`\gamma _A`$ and hence an $`(n3)`$-form on $`LM`$, depending on the connection $`A`$.
Then, for any representation $`\rho `$ of $`G`$, the following object is well-defined and gauge invariant in any dimension:
(2.1)
$$h_{k,\rho }(\gamma ;A,B)=\underset{0<t_1<\mathrm{}<t_k<1}{}\mathrm{Tr}_\rho \left[B(\dot{\gamma }_A(t_1))\mathrm{}B(\dot{\gamma }_A(t_k))\mathrm{Hol}(\gamma ;A)\right].$$
If $`n=3`$, then $`A+\kappa B`$ is also a connection and (2.1) is equal to the $`k`$-th Taylor coefficient in the $`\kappa `$-expansion of $`\mathrm{Tr}_\rho \mathrm{Hol}(\gamma ,A+\kappa B)`$. When $`n>3`$, then (2.1) is a differential form over $`LM`$ of degree $`k(n3)`$.
Unfortunately, however, (2.1) is not a good observable for the $`BF`$ theory since it is not invariant under the full set of its symmetries, see below (2.3).
If we modify (2.1) so as to get a โgood observable,โ then the relevant vacuum expectation values will produce elements of the $`k(n3)`$-rd cohomology of knots (imbedded loops) as explained in the following. Observe that in the case $`n=3`$ we recover the usual Vassiliev knot invariants.
### 2.1. Action functional and symmetries of $`BF`$ theory
We write the action functional of the $`BF`$ theory under the following
###### Assumption 1.
We assume that the Lie algebra $`๐ค`$ of $`G`$ possesses a nondegenerate $`\mathrm{Ad}`$-invariant bilinear form $`,`$. (For example, if $`๐ค`$ is semisimple, we may take the Killing form.)
We extend this form to $`\mathrm{\Omega }^{}(M,\mathrm{ad}P)`$ in the usual way. Then we define
(2.2)
$$S:=_MB,F_A,$$
where $`F_A`$ is the curvature $`2`$-form of $`A`$.
The EulerโLagrange equations of motion read
$$F_A=0,\mathrm{d}_AB=0;$$
thus, classically, $`BF`$ theory in $`n`$ dimensions describes flat connections together with covariantly closed $`(n2)`$-forms. Observe that โon shellโ (i.e., on the subspace of solutions) the covariant derivative is a coboundary operator.
###### Assumption 2.
We will assume in the following that $`M`$ is a compact manifold, that $`P`$ is a trivial bundle and that there is a flat connection $`A_0`$ on $`P`$ such that all cohomology groups $`H_{\mathrm{d}_{A_0}}^{}(M,\mathrm{ad}P)`$ are trivial.
Moreover, by โon shellโ we will always mean โon the subspace of those $`(A_0,B_0)`$ which satisfy the equations of motion with $`A_0`$ of this kind.โ
The symmetries under which the action is invariant correspond to an action of the group $`๐ข_{\mathrm{Ad}}\mathrm{\Omega }^{n3}(M,\mathrm{ad}P)`$, where $`๐ข`$ is the group of automorphisms of $`P`$ (i.e., ordinary gauge transformations). Infinitesimally we have
(2.3)
$$\begin{array}{cc}\hfill \delta A& =\mathrm{d}_A\xi ,\hfill \\ \hfill \delta B& =[B,\xi ]+\mathrm{d}_A\chi ,\hfill \end{array}$$
with $`(\xi ,\chi )\mathrm{\Omega }^0(M,\mathrm{ad}P)_{\mathrm{ad}}\mathrm{\Omega }^{n3}(M,\mathrm{ad}P)`$.
Observe that on shell this symmetries are โreducible;โ i.e., a covariantly closed $`\chi `$ (which under our triviality assumption is of the form $`\chi =\mathrm{d}_{A_0}\sigma `$) acts trivially.
More precisely, each point $`(A_0,B_0)`$ in the subspace of solutions has an isotropy group consisting of all covariantly closed $`(n3)`$-forms of the $`\mathrm{ad}`$ type.
With our triviality assumption, the isotropy group at each point is then isomorphic to the group $`\mathrm{\Omega }^{n4}(M,\mathrm{ad}P)/\mathrm{d}_{A_0}\mathrm{\Omega }^{n5}(M,\mathrm{ad}P)`$.
Of course, if $`n>5`$, also in this quotient there are nontrivial isotropy groups which are all isomorphic to $`\mathrm{\Omega }^{n5}(M,\mathrm{ad}P)/\mathrm{d}_{A_0}\mathrm{\Omega }^{n6}(M,\mathrm{ad}P)`$, and so on until we arrive at $`\mathrm{\Omega }^0(M,\mathrm{ad}P)`$ which acts freely on $`\mathrm{\Omega }^1(M,\mathrm{ad}P)`$.
In order to consistently gauge-fix all the symmetries, one has then to resort to the extended BRST formalism (i.e., introduce a hierarchy of ghosts for ghosts).
An additional problem is due to the fact that the isotropy groups are different off shell. So, in order to work in the Lagrangian formalism which is better suited for the perturbative expansions, one has to rely on the whole BatalinโVilkovisky (BV) machinery as explained in Section 3.
It is known that the partition function of $`BF`$ theory is related to the analytic torsion of $`M`$ (see for the abelian case and for the non-abelian one).
In order to get other topological invariants, one has to find interesting BV closed observables. In the rest of this paper we will discuss some of them, leaving most of the technical details to .
Before starting with the general discussion we recall the $`3`$-dimensional case as studied in .
### 2.2. The $`3`$-dimensional $`BF`$ theory
Since $`B`$ is a $`1`$-form now, one can add to the pure $`BF`$ action the so-called cosmological term,
$$S_3:=\frac{1}{6}_MB,[B,B],$$
and define
$$S_\kappa :=S+\kappa ^2S_3,\kappa .$$
This action is actually equal to the difference of two ChernโSimons actions evaluated at the connections $`A+\kappa B`$ and $`A\kappa B`$. From this observation one immediately gets the infinitesimal symmetries for this theory as
(2.4)
$$\begin{array}{cc}\hfill \delta _\kappa A& =\mathrm{d}_A\xi +\kappa ^2[B,\chi ],\hfill \\ \hfill \delta _\kappa B& =[B,\xi ]+\mathrm{d}_A\chi .\hfill \end{array}$$
Another consequence is that, for any loop $`\gamma `$ and any representation $`\rho `$ of the Lie algebra $`๐ค`$ of $`G`$, the Wilson loops $`\mathrm{Tr}_\rho \mathrm{Hol}(\gamma ;A\pm \kappa B)`$ are observables. We then have
(2.5)
$$\begin{array}{c}_\rho (\kappa ;\gamma ;A,B)=\mathrm{Tr}_\rho \mathrm{Hol}(\gamma ;A+\kappa B)=\mathrm{Tr}_\rho \mathrm{Hol}(\gamma ;A)+\underset{k=1}{\overset{\mathrm{}}{}}\kappa ^kh_{k,\rho }(\gamma ;A,B),\hfill \end{array}$$
where $`h_{k,\rho }(\gamma ;A,B)`$ is the $`k`$-th Taylor coefficient in the $`\kappa `$-expansion of $`\mathrm{Tr}_\rho \mathrm{Hol}(\gamma ;A+\kappa B)`$ introduced in (2.1).
By the previous discussion it follows clearly that both the even and the odd part of $``$ are observables.
In order to give an explicit description of $`h_{k,\rho }`$, it is better to view $`\gamma `$ as a periodic mapping from $`[0,1]`$ to $`M`$. We denote by $`H(\gamma ;A)|_s^t`$ the group element determined, in a given trivialization, by the $`A`$-parallel transport along $`\gamma `$ from the point $`\gamma (t)`$ to the point $`\gamma (s)`$. Then we can rewrite (2.1) in the form
$$\begin{array}{c}h_{k,\rho }(\gamma ;A,B)=_\mathrm{}_k\mathrm{Tr}_\rho [H(\gamma ;A)|_0^{t_1}B_1H(\gamma ;A)|_{t_1}^{t_2}B_2\mathrm{}\hfill \\ \hfill \mathrm{}B_{k1}H(\gamma ;A)|_{t_{k1}}^{t_k}B_kH(\gamma ;A)|_{t_k}^1],\end{array}$$
where $`\mathrm{}_k`$ is the $`k`$-simplex $`\{0<t_1<t_2<\mathrm{}<t_k<1\}`$ and $`B_i`$ is a shorthand notation for the pullback of $`B`$ via the map $`\gamma _i:\mathrm{}_kM`$, $`\gamma _i(t_1,\mathrm{},t_k)=\gamma (t_i)`$.
Observe that each $`h_{k,\rho }(;A,B)`$ defines a function on the loop space $`\mathrm{L}M`$ of $`M`$. It is not difficult to check thatโmodulo the equations of motion
$$F_A+\frac{\kappa ^2}{2}[B,B]=0,\mathrm{d}_AB=0,$$
of $`S_\kappa `$โthe functions $`(;A,B)`$ are locally constant on $`\mathrm{L}M`$.
Quantization then requires a regularization, viz., point splitting or, in a more precise formulation, the blowing up of the diagonals of configuration spaces. So one has to consider imbeddings, instead of generic loops, and to introduce a framing. The expectation values of the $``$s then define locally closed functions on the space $`\mathrm{Imb}_f(S^1,M)`$ of framed imbeddings, i.e., framed knot invariants.
### 2.3. A first glimpse to higher dimensions
A straightforward generalization of (2.5) to higher dimensions exists as discussed at the beginning of this section. It yields forms on $`\mathrm{L}M`$ instead of functions since only one form degree for each field $`B`$ is saturated by the integration. So now
$$h_{k,\rho }(;A,B)\mathrm{\Omega }^{(n3)k}(\mathrm{L}M).$$
Assume now that $`A`$ is a flat connection and that $`B`$ is covariantly closed, as classical solutions of the pure $`BF`$ theory are. Then it is not difficult to check that, for any odd $`n>3`$, $`h_{k,\rho }`$ is closed. This follows from the generalized Stokes theorem.
More precisely, $`h_{k,\rho }`$ is a form on $`\widehat{M}=\mathrm{L}M\times ๐\times \mathrm{\Omega }^{n2}(M,\mathrm{ad}P)`$, where $`๐`$ is the space of connections on $`M`$, and we restrict it to the subspace where $`F_A=\mathrm{d}_AB=0`$.
Let $`\widehat{\mathrm{d}}`$ be the differential on $`\widehat{M}`$. In the computation of $`\widehat{\mathrm{d}}h_{k,\rho }`$ we can switch the integral with the differential. We then get $`\widehat{\mathrm{d}}`$ acting on a function $`\eta _{k,\rho }`$ on $`\widehat{M}\times \mathrm{}_k`$:
$$\widehat{\mathrm{d}}h_{k,\rho }=_\mathrm{}_k\widehat{\mathrm{d}}\eta _{k,\rho }.$$
Let $`\mathrm{d}_k`$ be the differential on $`\mathrm{}_k`$ and $`\mathrm{d}:=\widehat{\mathrm{d}}\pm \mathrm{d}_k`$ the differential on $`\widehat{M}\times \mathrm{}_k`$. By adding and subtracting $`\mathrm{d}_k`$ we get then
$$\widehat{\mathrm{d}}h_{k,\rho }=_\mathrm{}_kd\eta _{k,\rho }_\mathrm{}_k\mathrm{d}_k\eta _{k,\rho }.$$
The first term on the r.h.s. is then easily seen to vanish in our hypotheses $`F_A=\mathrm{d}_AB=0`$ since
$$\mathrm{d}H(\gamma ;A)|_s^t=A(s)H(\gamma ;A)|_s^t+H(\gamma ;A)|_s^tA(t)$$
when $`A`$ is flat.
The second term can then be computed using the Stokes theorem. The codimension-one boundaries of the simplex corresponding to the collapse of consecutive points yield terms containing $`B^2`$ which vanishes for dimensional reasons ($`n>4`$). The remaining codimension-one boundaries correspond to $`t_0=0`$ or $`t_k=1`$. It is not difficult to check, using the cyclic property of the trace, that these terms cancel each other.
If the dimension $`n`$ is even, $`n>4`$, the only problem in the above discussion arises at the last step since, for $`k`$ even, the two terms coming from $`t_0=0`$ and $`t_k=1`$ sum up instead of canceling each other. For $`k`$ odd however everything works as before.
Similar computations allow to show that the $`h_{k,\rho }`$ are invariant (modulo exact forms on $`\mathrm{L}M`$) under the symmetries (2.3) either if $`n`$ is odd and greater then 5 or if $`n`$ is even and greater than 4 and $`k`$ is odd.
In sections 4 and 5 we will describe how this discussion can be extended โoff shellโ and how the cases $`n=4`$ and $`n=5`$ will be included.
## 3. The BV quantization of $`BF`$ theories
### 3.1. The BRST operator
In order to deal with the symmetries (2.3) of (2.2) in the functional integral, one has to introduce the BRST operator
(3.1)
$$\begin{array}{cc}\hfill \delta _{\mathrm{BRST}}A& =\mathrm{d}_Ac,\hfill \\ \hfill \delta _{\mathrm{BRST}}B& =[B,c]+\mathrm{d}_A\tau _1,\hfill \end{array}$$
where $`c`$ and $`\tau _1`$ are ghosts, i.e., forms on the space of fields with values in $`\mathrm{\Omega }^0(M,\mathrm{ad}P)`$ and $`\mathrm{\Omega }^{n3}(M,\mathrm{ad}P)`$ respectively. As usual in gauge theories one also defines
$$\delta _{\mathrm{BRST}}c=\frac{1}{2}[c,c].$$
Because of the on-shell reducibility, one has then to introduce ghosts for ghosts $`\tau _k`$ with values in $`\mathrm{\Omega }^{n2k}(M,\mathrm{ad}P)`$, $`k=1,\mathrm{},n2`$, with ghost number equal to $`kmod2`$ and extended BRST operator
$$\begin{array}{cc}\hfill \delta _{\mathrm{BRST}}\tau _k& =(1)^k[\tau _k,c]+\mathrm{d}_A\tau _{k+1},k=1,\mathrm{},n3,\hfill \\ \hfill \delta _{\mathrm{BRST}}\tau _{n2}& =(1)^n[\tau _{n2},c].\hfill \end{array}$$
### 3.2. The BV formalism
It is not difficult to check that $`\delta _{\mathrm{BRST}}^2=0modF_A`$.
The BatalinโVilkovisky method allows then for the construction of a nilpotent operator $`\delta _{\mathrm{BV}}`$ that extends $`\delta _{\mathrm{BRST}}`$ off shell.
To do so, one first introduces a partner $`\varphi _\alpha ^+`$ with values in $`\mathrm{\Omega }^{}(M,\mathrm{ad}P)`$ for any field or ghost $`\varphi ^\alpha =A,B,c,\tau _1,\mathrm{},\tau _{n2}`$ with the following rules:
* The ghost number of $`\varphi _\alpha ^+`$ is minus the ghost number of $`\varphi ^\alpha `$, minus one.
* The form degree of $`\varphi _\alpha ^+`$ is $`n`$ minus the form degree of $`\varphi ^\alpha `$.
###### Remark 3.1.
In general, the antifields are dual to the corresponding fields. Of course some isomorphisms may be used to identify certain spaces. For example here we have preferred to identify the Lie algebra $`๐ค`$ with its dual (using the bilinear from $`,`$) so that also the antifields take values in the space of tensorial forms of the $`\mathrm{ad}`$ type. This will be particularly useful, e.g., in equations (3.5) and (3.6).
In the original formulation of Batalin and Vilkovisky one also identifies forms of complementary degree using a Hodge operator. Since we do not want to introduce a metric here, we prefer to avoid this identification. As a consequence, our BV antibracket (3.2) will be of the form described in instead of the original one.
###### Remark 3.2 (Sign convention).
We follow here the usual convention for the sign rules related to the double grading given by the form degree $`\mathrm{deg}`$ and the ghost number $`\mathrm{gh}`$.
Namely, in the case of homogenous forms $`\alpha `$ and $`\beta `$ of the $`\mathrm{ad}`$-type we have
$$[\alpha ,\beta ]=(1)^{\mathrm{deg}\alpha \mathrm{deg}\beta +\mathrm{gh}\alpha \mathrm{gh}\beta }[\beta ,\alpha ].$$
Moreover, in the case of homogenous forms $`\alpha `$ and $`\beta `$ taking values in a commutative algebra (e.g., $``$) we have
$$\alpha \beta =(1)^{\mathrm{deg}\alpha \mathrm{deg}\beta +\mathrm{gh}\alpha \mathrm{gh}\beta }\beta \alpha .$$
Next we define the BV bracket of two functionals $`F`$ and $`G`$. We use throughout Einsteinโs convention over repeated indices and set
(3.2)
$$(F,G):=_MF\frac{\stackrel{}{}}{\varphi ^\alpha },\frac{\stackrel{}{}}{\varphi _\alpha ^+}G(1)^{\mathrm{deg}\varphi ^\alpha (n+1)}F\frac{\stackrel{}{}}{\varphi _\alpha ^+},\frac{\stackrel{}{}}{\varphi ^\alpha }G,$$
where the left and right functional derivatives are given by the following formula:
$$\frac{\mathrm{d}}{\mathrm{d}t}|_{t=0}F(\varphi ^\alpha +t\rho ^\alpha )=_M\rho ^\alpha ,\frac{\stackrel{}{}}{\varphi ^\alpha }F=_MF\frac{\stackrel{}{}}{\varphi ^\alpha },\rho ^\alpha ;$$
we proceed similarly for the antifields.
As usual, the space of functionals with BV bracket is a Gerstenhaber algebra .
Finally the BV operator is defined by
(3.3)
$$\delta _{\mathrm{BV}}:=(S_{\mathrm{BV}},)$$
where $`S_{\mathrm{BV}}`$ is a solution of the master equation
$$(S_{\mathrm{BV}},S_{\mathrm{BV}})=0$$
such that $`S_{\mathrm{BV}}^{}{}_{|_{\varphi _{}^+=0}}{}^{}=S`$.
### 3.3. The BV action for $`BF`$ theories
The BV action $`S_{\mathrm{BV}}`$ corresponding to the $`BF`$ action (2.2) can be written as
(3.4)
$$S_{\mathrm{BV}}=_M๐ก;๐ฅ_๐ $$
where the notations are as follows:
* The dot product is just the wedge product between forms taking values in an associative algebraโe.g., $``$ or $`๐ค`$ itself if it associative as in Section 5โbut with a shifted degree; viz., for two homogenous forms $`\alpha `$ and $`\beta `$ we set
$$\alpha \beta :=(1)^{\mathrm{gh}\alpha \mathrm{deg}\beta }\alpha \beta ,$$
where $`\mathrm{gh}`$ denotes the ghost number.
We extend then the bilinear form $`,`$ to forms with shifted degree by setting
$$\alpha ;\beta :=(1)^{\mathrm{gh}\alpha \mathrm{deg}\beta }\alpha ,\beta .$$
Similarly we define the dot Lie bracket for two homogeneous forms of the $`\mathrm{ad}`$ type by
$$[[\alpha ;\beta ]]:=(1)^{\mathrm{gh}\alpha \mathrm{deg}\beta }[\alpha ,\beta ].$$
These definitions are then extended by linearity.
An easy check shows that the dot product in the case of a commutative algebra and the dot Lie bracket are, respectively, a graded commutative product and a graded Lie bracket with respect to a new grading called the total degree that is defined as the form degree plus the ghost number. Moreover, $`\mathrm{d}_{A_0}`$ is still a differential for the dot algebras.
* The โsuper $`B`$-fieldโ $`๐ก`$ is defined by
(3.5)
$$๐ก=\underset{k=1}{\overset{n2}{}}(1)^{\frac{k(k1)}{2}}\tau _k+B+(1)^nA^++c^+\mathrm{\Omega }^{}(M,\mathrm{ad}P),$$
and has total degree equal to $`n2`$.
* The โsupercurvatureโ $`๐ฅ_๐ `$ of the โsuperconnectionโ
(3.6)
$$๐ =(1)^{n+1}c+A+(1)^nB^++\underset{k=1}{\overset{n2}{}}(1)^{n(k+1)+\frac{k(k1)}{2}}\tau _k^+$$
is given by the usual formula. In order to write it down, it is better to choose a background connection $`A_0`$ and to define the tensorial form $`๐บ=๐ A_0`$ of total degree one. Then
$$๐ฅ_๐ =F_{A_0}+\mathrm{d}_{A_0}๐บ+\frac{1}{2}[[๐บ;๐บ]].$$
In general we will choose $`A_0`$ to be a flat connection as in Assumption 2.
* By $`_M`$ we then mean the integral of all the terms of form degree equal to $`n`$. Observe that as a consequence $`S_{\mathrm{BV}}`$ has then ghost number zero.
###### Remark 3.3.
We may observe that there is a superspace formulation of (3.4) obtained by introducing superpartners to the coordinates of $`M`$ and redefining $`๐ `$ and $`๐ก`$ accordingly. In this way we would follow the pattern described in .
Special cases were already discussed in (two dimensions) and (four dimensions). See also for the case of the $`3`$-dimensional ChernโSimons theory.
For later purposesโviz., in order to define loop observables as in the following sectionsโit is however better to work in our setting.
We have the following general result :
###### Theorem 1.
The action $`S_{\mathrm{BV}}`$ satisfies the master equation in any dimension.
We conclude this section by giving the explicit action of the BV operator (3.3) on the โsuperfieldsโ $`๐ `$ and $`๐ก`$. In order to give neater formulae, it is better to define a new BV operator with shifted degree; viz., for a homogeneous form $`\alpha `$, we set
$$๐น\alpha :=(1)^{\mathrm{deg}\alpha }\delta _{\mathrm{BV}}\alpha .$$
One can show that $`๐น`$ is a differential for the dot algebras and that it anticommutes with $`\mathrm{d}_{A_0}`$.
Then we obtain
(3.7)
$$\begin{array}{cc}\hfill ๐น๐ & =(1)^n๐ฅ_๐ ,\hfill \\ \hfill ๐น๐ก& =(1)^n\mathrm{d}_๐ ๐ก,\hfill \end{array}$$
with
$$\mathrm{d}_๐ ๐ก=\mathrm{d}_{A_0}๐ก+[[๐บ;๐ก]].$$
Upon using the above equations, we can then prove Thm. 1 by simply checking that $`๐นS_{\mathrm{BV}}=0`$, as follows from the the $`\mathrm{ad}`$-invariance of $`,`$ and from the Stokes theorem.
### 3.4. The BV Laplace operator and the BV observables
In the quantum version of the BV formalismโi.e., when dealing with functional integrals with weight $`\mathrm{exp}(\mathrm{i}/\mathrm{})S`$โone has then to introduce the so-called BV Laplace operator $`\mathrm{\Delta }_{\mathrm{BV}}`$ and to verify that the quantum master equation
$$(S_{\mathrm{BV}},S_{\mathrm{BV}})2\mathrm{i}\mathrm{}\mathrm{\Delta }_{\mathrm{BV}}S_{\mathrm{BV}}=0$$
is satisfied.
The very definition of $`\mathrm{\Delta }_{\mathrm{BV}}`$ relies on a regularization of the theory, which we do not discuss here. We only recall the formal properties of $`\mathrm{\Delta }_{\mathrm{BV}}`$; viz.:
1. $`\mathrm{\Delta }_{\mathrm{BV}}`$ is a coboundary operator on the space of functionals;
2. for any two functionals $`F`$ and $`G`$,
(3.8)
$$\mathrm{\Delta }_{\mathrm{BV}}(FG)=(\mathrm{\Delta }_{\mathrm{BV}}F)G+(1)^{\mathrm{gh}F}F\mathrm{\Delta }_{\mathrm{BV}}G+(1)^{\mathrm{gh}F}(F,G).$$
The space of functionals with the BV bracket and the BV Laplacian is a so-called BV algebra.
To give an explicit definition of the BV Laplacian one has to introduce some extra structures (e.g., a Riemannian metric on $`M`$) and a regularization. The main property however is that the BV Laplace operator contracts each field with the Hodge dual of the corresponding antifield at the same point in $`M`$.
Under our assumptions, one can then prove that $`\mathrm{\Delta }_{\mathrm{BV}}S_{\mathrm{BV}}=0`$ for $`S_{\mathrm{BV}}`$ in (3.4).
So $`S_{\mathrm{BV}}`$ is also a solution of the quantum master equation. This implies that its partition function is independent of the choice of gauge fixing.
A consequence of the properties of the BV operators is that the operator
$$\mathrm{\Omega }_{\mathrm{BV}}:=\delta _{\mathrm{BV}}\mathrm{i}\mathrm{}\mathrm{\Delta }_{\mathrm{BV}}$$
is a coboundary operator iff $`S_{\mathrm{BV}}`$ satisfies the quantum master equation.
The main statement in the BV formalism is that the $`\mathrm{\Omega }_{\mathrm{BV}}`$-cohomology of ghost number zero yields all the meaningful observables. More precisely, this means that the expectation value of an $`\mathrm{\Omega }_{\mathrm{BV}}`$-closed functional is independent of the gauge fixing and that the expectation value of an $`\mathrm{\Omega }_{\mathrm{BV}}`$-exact functional (or of a functional of ghost number different from zero) vanishes.
In the next sections we will discuss some BV observables of $`BF`$ theories associated to $`\mathrm{L}M`$ (or better to $`\mathrm{Imb}_f(S^1,M)`$). To do so, it is however better to use shifted versions of the operators $`\mathrm{\Omega }_{\mathrm{BV}}`$ and $`\mathrm{\Delta }_{\mathrm{BV}}`$ as well, viz.:
$`๐\alpha `$ $`:=(1)^{\mathrm{deg}\alpha }\mathrm{\Omega }_{\mathrm{BV}}\alpha ,`$
$`๐ซ\alpha `$ $`:=(1)^{\mathrm{deg}\alpha }\mathrm{\Delta }_{\mathrm{BV}}\alpha .`$
## 4. Generalized Wilson loops in odd dimensions
At this point we are ready to define the correct generalization of the observables $``$ defined in (2.5).
Formally the new observable is still the trace of the โholonomy of $`๐ +\kappa ๐ก`$
(4.1)
$$๐_\rho (\kappa ;๐ ,๐ก):=\mathrm{Tr}_\rho \mathrm{๐ง๐๐
}(๐ +\kappa ๐ก)\mathrm{\Omega }^{}(\mathrm{L}M),$$
where the โholonomyโ $`\mathrm{๐ง๐๐
}`$ is now defined in terms of iterated integrals as follows: First we write $`๐ =A_0+๐บ`$. Then we set
$$\mathrm{Tr}_\rho \mathrm{๐ง๐๐
}(๐ +\kappa B):=\mathrm{Tr}_\rho \mathrm{Hol}(A_0)+\underset{l=1}{\overset{\mathrm{}}{}}h_{l,\rho }(A_0,๐บ+\kappa ๐ก),$$
where
$$\begin{array}{c}h_{l,\rho }(A_0,๐บ+\kappa ๐ก)=_\mathrm{}_l\mathrm{Tr}_\rho [H(A_0)|_0^{t_1}(๐บ_1+\kappa ๐ก_1)H(A_0)|_{t_1}^{t_2}\hfill \\ \hfill (๐บ_2+\kappa ๐ก_2)\mathrm{}H(A_0)|_{t_{l1}}^{t_l}(๐บ_l+\kappa ๐ก_l)H(A_0)|_{t_l}^1].\end{array}$$
Here $`๐บ_i`$ and $`๐ก_i`$ are shorthand notations for the pullbacks of $`๐บ`$ and $`๐ก`$ via $`\mathrm{ev}_i:\mathrm{L}M\times \mathrm{}_lM`$, $`(\gamma ;t_1,\mathrm{},t_l)\gamma (t_i)`$. Moreover, the $`H(A_0)|_s^t`$โs denote the group elements associated to parallel transports as functions on $`\mathrm{L}M`$.
###### Remark 4.1.
The above integrals should be better viewed as integrations along the fiber of the trivial bundles $`\mathrm{}_l\times \mathrm{L}M\mathrm{L}M`$. That is, the integrals are zero whenever the form degree is less than the dimension of the simplex and yield a form on $`\mathrm{L}M`$ whenever the form degree exceeds the dimension of the simplex.
Also observe that $`๐`$ is a sum of terms with different ghost number and different form degree on $`\mathrm{L}M`$.
Of course, we cannot expect $`๐`$ to be an observable for the pure $`BF`$ theory. We can however consider a fake โhigher dimensional $`BF`$ theory with cosmological termโ as follows: We first define
(4.2)
$$๐ฒ_3(๐ก):=\frac{1}{6}_M๐ก;[[๐ก;๐ก]],$$
where again we consider only the terms of form degree equal to $`n`$, so $`๐ฒ_3`$ has ghost number $`2(n3)`$. Then we consider the functional
(4.3)
$$\widehat{๐}_\rho (\mathrm{},\kappa ;๐ ,๐ก):=\left\{\mathrm{exp}[(\mathrm{i}/\mathrm{})\kappa ^2๐ฒ_3(๐ก)]๐_\rho (\kappa ;๐ ,๐ก)\right\}_0,$$
where $`\{\}_0`$ means taking the terms with ghost number zero.
The functional $`\widehat{๐}_\rho `$ is well-defined for any loop in $`M`$. However, in order to avoid problems with the BV Laplace operator $`\mathrm{\Delta }_{\mathrm{BV}}`$, we must restrict ourselves to the space of framed imbeddings $`\mathrm{Imb}_f(S^1,M)`$.
###### Theorem 2.
For any $`\kappa `$ and $`\rho `$ and any odd dimension $`n`$, $`\widehat{๐}`$, as a functional taking values in the forms on $`\mathrm{Imb}_f(S^1,M)`$, is $`\mathrm{\Omega }_{\mathrm{BV}}`$-closed modulo $`\mathrm{d}`$-exact forms and $`\mathrm{d}`$-closed modulo $`\mathrm{\Omega }_{\mathrm{BV}}`$-exact terms. In other words, $`[\widehat{๐}]`$ is an $`H^{}(\mathrm{Imb}_f(S^1,M))`$-valued observable.
###### Proof (Sketch).
The main idea of the proof relies on the identity
(4.4)
$$\begin{array}{c}๐\left\{\mathrm{exp}[(\mathrm{i}/\mathrm{})\kappa ^2๐ฒ_3(๐ก)]๐_\rho (\kappa ;๐ ,๐ก)\right\}=\mathrm{exp}[(\mathrm{i}/\mathrm{})\kappa ^2๐ฒ_3(๐ก)]๐น_\kappa ๐_\rho (\kappa ;๐ ,๐ก),\hfill \end{array}$$
where $`๐น_\kappa `$ is the following coboundary operator:
$$\begin{array}{cc}\hfill ๐น_\kappa ๐ & =๐ฅ_๐ \frac{\kappa ^2}{2}[[๐ก;๐ก]],\hfill \\ \hfill ๐น_\kappa ๐ก& =\mathrm{d}_๐ ๐ก.\hfill \end{array}$$
Observe that for $`n3`$, $`๐น_\kappa `$ is a differential only for the $`_2`$-reduction of the graded algebra of functionals.
Using $`๐น_\kappa `$ is like working with a cosmological term, and, upon using the generalized Stokes theorem, one gets
$$(\mathrm{d}+๐น_\kappa )๐_\rho (\kappa ;๐ ,๐ก)=0,$$
which proves the theorem.
Equation (4.4) is a consequence of (3.8) and of the following identities:
$$๐น๐ฒ_3=0,๐ซ\mathrm{exp}[(\mathrm{i}/\mathrm{})\kappa ^2๐ฒ_3(๐ก)]=0,๐ซ๐_\rho =0.$$
The first identity follows from (3.7), from the fact that $`,`$ is $`\mathrm{ad}`$-invariant and from Stokes theorem.
The second identity holds since $`๐ฒ_3`$ depends only on $`๐ก`$ and as a consequence of the already discussed property according to which the BV Laplace operator contracts each field with the Hodge dual (for some Riemannian metric on $`M`$) of the corresponding antifield at the same point in $`M`$.
The last identity is โformallyโ (i.e., modulo regularization problems for $`\mathrm{\Delta }_{\mathrm{BV}}`$) true if $`\gamma `$ does not have transversal self-intersections, for the same reason as above. However, in order to rely upon this last identity confidently, we must then restrict ourselves to framed imbeddings and put each component of $`๐ `$ on the imbedding and each component of $`๐ก`$ on its companion (as done in ). โ
As a consequence, the expectation value of $`\widehat{๐}`$ is (up to anomalies) a cohomology class on the space of (framed) imbeddings of $`S^1`$ into $`M`$.
###### Remark 4.2.
If we set all the antifields to zero, $`\widehat{๐}`$ reduces to a sum of $`h_{k,\rho }(A,B)`$โs; so it is the off-shell generalization we were looking for in subsection 2.3.
Moreover, the expectation value of $`\widehat{๐}`$ w.r.t. the pure $`BF`$ theory is in three dimensions the same as the expectation value of $``$ in the $`BF`$ theory with cosmological term.
###### Remark 4.3.
$`\widehat{๐}`$ is a genuine quantum observable since its limit for $`\mathrm{}0`$ is not defined.
However, one might replace $`\kappa `$ with $`\mathrm{}\kappa `$. In this way, $`\widehat{๐}`$ becomes a formal power series in $`\mathrm{}`$. The zeroth order term is just $`\mathrm{Tr}_\rho \mathrm{Hol}(A)`$.
Since $`\widehat{๐}`$ is an observable for any $`\kappa `$, so is its odd part (in $`\kappa `$) $`\widehat{๐}^o`$. It is not difficult to see that $`\widehat{๐}_\rho ^o(\mathrm{},\mathrm{}\kappa ;๐ ,๐ก)/\mathrm{}`$ is as well a formal power series in $`\mathrm{}`$ and that its zeroth-order term is the observable
$$๐_{1,\rho }(๐ ,๐ก):=\frac{\mathrm{d}}{\mathrm{d}\kappa }|_{\kappa =0}h_{1,\rho }(A_0,๐บ+\kappa B),$$
which is the off-shell extension of $`h_{1,\rho }(A,B)`$.
Therefore, the observables $`\widehat{๐}_\rho (\mathrm{},\mathrm{}\kappa ;๐ ,๐ก)`$ and $`\widehat{๐}_\rho ^o(\mathrm{},\mathrm{}\kappa ;๐ ,๐ก)/\mathrm{}`$ are nontrivial quantum deformations of, respectively, the ordinary Wilson loop and $`๐_{1,\rho }`$.
## 5. Other loop observables
In order to generalize some of the results of the previous section and in order to define more general observables we make the following
###### Assumption 3.
We assume that the Lie algebra $`๐ค`$ is obtained from an associative algebra with trace $`\mathrm{Tr}`$ (e.g., we may take $`๐ค=๐ค๐ฉ(N)`$ with the usual trace of matrices). In this case, we assume that our $`\mathrm{ad}`$-invariant bilinear form $`\xi ,\eta `$ is given by $`\mathrm{Tr}(\xi \eta )`$ and, according to Assumption 1, we further assume that it is nondegenerate. Moreover, we consider only representations $`\rho `$ of $`๐ค`$ as an associative algebra.
### 5.1. Generalized Wilson loops in even dimensions
The observable defined in the previous section does not work in even dimensions essentially because the โcosmological termโ $`๐ฒ_3`$ vanishes when $`๐ก`$ has even total degree.
We can cure this problem thanks to Assumption 3 by defining instead
$$๐ฎ_3(๐ก):=\frac{1}{3}_M\mathrm{Tr}(๐ก๐ก๐ก).$$
We have already seen in subsection 2.3 that the even part of $``$ does not work. So we consider only
(5.1)
$$\widehat{๐}_\rho ^o(\mathrm{},\kappa ;๐ ,๐ก):=\left\{\mathrm{exp}[(\mathrm{i}/\mathrm{})\kappa ^2๐ฎ_3(๐ก)]๐_\rho ^o(\kappa ;๐ ,๐ก)\right\}_0,$$
where $`๐^o`$ is the odd part of $`๐`$, which is defined exactly as in the odd-dimensional caseโsee equations (4.1) and following.
We have then the following analogue (with analogous proof) of Thm. 2:
###### Theorem 3.
For any $`\kappa `$ and $`\rho `$ and any even dimension $`n`$, $`[\widehat{๐}^o]`$ is an $`H^{}(\mathrm{Imb}_f(S^1,M))`$-valued observable.
###### Remark 5.1.
Similarly to what happens in the odd-dimensional case, $`\widehat{๐}^o`$ reduces to a sum of $`h_{2k+1,\rho }`$ as the antifields are set to zero.
Moreover, $`\widehat{๐}_\rho ^o(\mathrm{},\mathrm{}\kappa ;๐ ,๐ก)/\mathrm{}`$ is still a nontrivial quantum deformation of $`๐_{1,\rho }`$.
However, we do not find in even dimensions a nontrivial quantum deformation of the ordinary Wilson loop $`\mathrm{Tr}_\rho \mathrm{Hol}`$.
### 5.2. Loop observables with more then cubic interactions
The โcosmological termsโ $`๐ฒ_3`$ and $`๐ฎ_3`$ give rise, in the perturbative expansion, to trivalent vertices.
If we work with Assumption 3, we can define more general interaction terms:
$$๐ฎ_r(๐ก):=\frac{1}{r}_M\mathrm{Tr}๐ก^r.$$
Observe that in odd dimensions $`๐ฎ_r`$ vanishes if $`r`$ is even.
Next, for any two given sequences $`๐=\{\mu _1,\mu _2,\mathrm{}\}`$ and $`๐=\{\lambda _1,\lambda _2,\mathrm{}\}`$, we define
(5.2)
$$\begin{array}{c}\stackrel{~}{๐}_\rho (\mathrm{},๐,๐;๐ ,๐ก):=\left\{\mathrm{exp}\left[(\mathrm{i}/\mathrm{})\underset{r=1}{\overset{\mathrm{}}{}}\mu _r๐ฎ_{r+1}(๐ก)\right]\mathrm{Tr}_\rho \mathrm{๐ง๐๐
}\left(๐ +\underset{s=1}{\overset{\mathrm{}}{}}\lambda _s๐ก^s\right)\right\}_0.\hfill \end{array}$$
We then denote by $`\stackrel{~}{๐}_\rho ^o`$ the odd part of $`\stackrel{~}{๐}_\rho `$ under $`๐๐`$.
We have then
###### Theorem 4.
In odd dimensions, $`[\stackrel{~}{๐}_\rho ]`$ is an $`H^{}(\mathrm{Imb}_f(S^1,M))`$-valued observable whenever the following conditions are satisfied
$$\mu _{2l1}=\lambda _{2l}=0,l,$$
$$\mu _{2l}=\underset{\begin{array}{c}i,j0\\ i+j=l1\end{array}}{}\lambda _{2i+1}\lambda _{2j+1},l.$$
In even dimensions, $`[\stackrel{~}{๐}_\rho ^o]`$ is an $`H^{}(\mathrm{Imb}_f(S^1,M))`$-valued observable whenever the following conditions are satisfied
$$\mu _l=\underset{\begin{array}{c}i,j1\\ i+j=l\end{array}}{}\lambda _i\lambda _j,l.$$
The proof is a direct generalization of the proof of Thm. 2.
## 6. Conclusions
In this paper we have defined some $`H^{}(\mathrm{Imb}_f(S^1,M))`$-valued observables for $`BF`$ theories on a trivial principal $`G`$-bundle $`PM`$ associated to any representation of the Lie algebra $`๐ค`$.
Our ideas extend naturally to nontrivial bundles as well, though we did not consider this extension here for the sake of simplicity.
The expectation values of these observables define then classes in the cohomology of the space of framed imbeddings of $`S^1`$ into $`M`$, which we have assumed to be compact in order to simplify the discussion.
Of course a very interesting case is $`M=^n`$, which is not compact. The only extra technical point here is that one has to specify the correct behavior at infinity of all the fields and antifields. This done, the trivial connection satisfies the hypotheses of Assumption 2.
The perturbative expansion of the expectation values in the case $`M=^n`$ around the trivial connection is then obtained in terms of the configuration space integrals discussed in , where however only the framing-independent coholomogy classes were considered explicitly.
Notice that in this paper we have not defined observables with trivalent interactions and an even number of $`B`$-fields placed on the imbedding in the even-dimensional case. Thus, we cannot obtain the nontrivial class of imbeddings represented by the diagram cocycle of Figure 4 in . This suggests that there might exist other observables than those we have considered here.
On the other hand, one may use the combinatorics of this quantum field theory to obtain new nontrivial diagram cocycles. |
warning/0003/cond-mat0003369.html | ar5iv | text | # Unusual magnetic relaxation behavior in La0.5Ca0.5MnO3
## Abstract
We present an extensive study of the time dependence of the magnetization in a polycrystalline and low temperature charge ordered La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> sample. After application and removal of a 5 T magnetic field, a systematic variation of the magnetic relaxation rate from 10 K to 245 K was found. At 195 K, the magnetization decreases in a very short time and after that it increases slowly as a function of time. Moreover, between 200 and 245 K, an increase in magnetization, above the corresponding value just after removing the 5 T magnetic field, was measured. This unusual behavior was tested in several other relaxation procedures.
PACS: 70, 74.25 Ha, 75.60.-d, 76.60.Es
The low temperature charge ordered compound La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> has been intensively studied in the last years due to its very large variations in resistivity, magnetization, and lattice parameters as a function of temperature, magnetic field and isotope mass<sup>,</sup><sup>,</sup>. Besides, this compound is particularly interesting due to the coexistence and microscopic separation of two phases at low temperatures: one ferromagnetic and one antiferromagnetic. Recently, Huang et. al., measured neutron powder diffraction and magnetization in La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> samples, with x=0.47, 0.50 and 0.53. They argued the existence in all samples of a paramagnetic-ferromagnetic transition, for a phase called F-I, at 265 K and the formation of a second crystallographic phase, named A-II, below 230 K. They also pointed out that phase A-II ordered antiferromagnetically with a CE-type magnetic structure below 160 K. Moreover, Radaelli et. al. measured a rapid change of the lattice parameters between 130 K and 225 K in La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub>. This was associated with the development of a Jahn-Teller distortion of the Mn-O octahedra, as well as partial orbital ordering.
Relaxation experiments are a useful tool to study the dynamics of the charge ordering phase due to the frustration in the spin equilibrium configuration. Until now, attention has been focused in the relaxation of electrical resistivity after a large change of an applied magnetic field, which induces a transition from a ferromagnetic metallic state to a charge ordered insulator phase or conversely<sup>,</sup>. However, systematic reports of magnetic relaxation curves (M(t)) in charge ordered compounds are rare, possible because the abrupt jump seen in resistivity is absent in magnetic measurements. Here, we present an extensive study M(t) curves measured in a polycrystalline sample of La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub>. We also performed similar measurements in a polycrystalline La<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> sample for comparison.
Polycrystalline samples of La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> and La<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> were prepared by the standard procedures described elsewhere <sup>,</sup>. X-ray diffraction measurements pointed out high quality samples. Magnetization measurements were done with a standard MPMS-5S SQUID magnetometer. The relaxation measuring procedure was the following: first, the sample was heated to 400 K in zero magnetic field; second, the remanent magnetic field in the solenoid of the SQUID magnetometer was seted to zero; third, the sample was cooled down to the working temperature in zero magnetic field; fourth, an applied magnetic field (H) was increased from 0 to 5 T at a rate of 0.83 T/minute and remained applied for a waiting time t<sub>w</sub>=50 s; fifth, H was decreased to zero at the same rate; finally, when H was zero (we defined this moment as t=0) the M(t) curve was recorded for more than 210 minutes. We have measured the profile of the remanent magnetic field trapped in the superconducting solenoid after increasing H to 5 T and the subsequent removal. Within the experimental region the trapped magnetic field was lower than 1.1 mT.
Figure 1 shows the temperature dependence of the magnetization measured with H=5 T for one of the La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> samples studied. The inset shows the same type of measurement with H=1.2 mT. In the first case the measurement started at 2 K after a zero field cooling procedure, while in the second (inset) started at 400 K. The large hysteresis, at magnetic field as high as 5 T, is a clear evidence of the intrinsic frustration in the equilibrium configuration of the spin system. The paramagnetic-ferromagnetic transition around 225 K (minimum slope) is independent of the applied field and the followed path. However, the transition to the antiferromagnetic phase is path dependent and the Neel temperature (T<sub>N</sub>) decreases with the increase of H. Values of T<sub>N</sub>, calculated from the maximum slope in the curves, changed between 100 and 195 K. The shape of these curves is in agreement with previous reports in the literature <sup>,</sup>.
Figure 2 shows magnetic relaxation measurements from 145 K to 195 K. To facilitate the comparison between curves at different temperatures, the magnetization in each case has been normalized to the corresponding value just after removing the H=5 T. We will denote these curves as m(t)=M(t)/M(0). As could be seen, the relaxation rate (mean slope of the curves or $`\mathrm{\Delta }`$m/$`\mathrm{\Delta }t`$) increases systematically with increasing temperatures from 150 K to 195 K. It is important to note that slopes here are negatives. The ratio of magnetization change between t=$`\mathrm{}`$ (the last measurement made) and t=0, m($`\mathrm{}`$), at 150 K is about 20 %, while at 195 K is only 0.9 %. The increase in the relaxation rate between 150 and 195 K is in clear contrast with the behavior between 10 K and 150 K, where the relaxation rate decreases. Included in figure 2 is the curve corresponding to 145 K which illustrates this behavior. Measurements of M versus H for this system <sup>,</sup> revealed the complete destruction of the antiferromagnetic phase above 150 K with H=5 T. This fact could be influencing the change of behavior of the relaxation rate around 150 K.
The inset in figure 2 reproduces the relaxation curve before normalization for 195 K. This curve is noticeable different from those at lower temperatures. Here, the magnetization first decreases and, after approximately 4 minutes, starts increasing with time. The inset in figure 3 also shows the M(t) curves at 200 K. In this case, no decrease in magnetization was measured, but a monotonous increase with time (approximately 62$``$10$`{}_{}{}^{5}\mu _{B}^{}`$ in 218 minutes) is observed. The experiment at 200 K was repeated with a sample of the same compound having only 9 % of the original mass. A similar increase of magnetization with time (88$``$10$`{}_{}{}^{5}\mu _{B}^{}`$ in 218 minutes) was found. These values are about 10 times smaller than the magnetization measured with H=1.2 mT at the same temperature (see inset in figure 1).
The main frame of figure 3 shows m(t) curves from 195 K to 245 K. Notice that, differently from the previous figure, all curves, except the one at 195 K, show values above one. In other words, the magnetization increases with time (curves have positive slopes) above the M(0) value in each case. Furthermore, m($`\mathrm{}`$) values are systematically higher with higher temperatures: from 0.9 % at 195 K to 75 % at 240 K. However, M(0) decreases with higher temperatures, as shown in figure 1. The curve at 245 K is also shown and presents a smaller m($`\mathrm{}`$) value with respect to the one at 240 K, probably associated with the transition of the system to the paramagnetic phase. Magnetic relaxation measurements were also done between 245 K and 350 K. In this temperature interval we did not find a systematic variation of the relaxation rate. Nonetheless, above 260 K, the M(t) curves always show the usual decreasing behavior with time.
As a small magnetic field is trapped (typically about 1 mT) in the superconducting solenoid of the SQUID magnetometer, resulting after successive applications and removals of H=5 T, we decided to check its role in the unusual increase in magnetization. The temperature of 210 K was chosen since the absolute value of magnetization is not too small and the increase in normalized magnetization is higher than at 200 K. We repeated the standard relaxation procedure, but instead of changing the applied magnetic field from 5 T to H=0 mT (squares), we reduced it from 5 T to H=-1.2 mT (circles) and H=-10 mT (triangles in the inset). In each case, this last field remained applied during the whole relaxation measurement. For H=-1.2 mT the actual field applied to the sample is almost zero. The initial magnetization of the sample here is lower than in the H=0 mT case, but it is still positive. However, the H=-10 mT is about 10 times higher than the trapped flux in the SQUID solenoid and the actual applied field to the sample is negative. The initial magnetization of the sample is lower with H=-10 mT than in the previous cases and has a negative value (see inset). Note that, contrary to what it is usually expected, the magnetization increases with time even with applied negative magnetic fields. The last measurements strongly support the point that is not the small trapped magnetic field in the solenoid of the magnetometer which is causing the unusual relaxation in La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub>.
Next, we repeated the standard relaxation procedure changing t<sub>w</sub>. Figure 5 shows three m(t) curves with t<sub>w</sub> equal to 50 s (squares), 500 s (circles) and 5000 s (up triangles). Clearly, the normalized increase in magnetization is higher the longer the 5 T magnetic field remains applied. Values of M(0) also increase for longer t<sub>w</sub>. A plausible explanation could be that the remanent trapped field in the sample after removing the H=5 T, which is higher for longer t<sub>w</sub>, is causing the self-alignment of the spins and the increase in magnetization. Therefore, these curves would be reflecting information about the interactions in the La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> sample.
Finally, we repeated the standard relaxation procedure for a La<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> polycrystalline sample. The ground state of this composition is a single ferromagnetic phase with the paramagnetic-ferromagnetic transition around 250 K. The magnetization, in relaxation measurements both above and below the Curie temperature, always decreased with time. These results for La<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> are in agreement with a similar study reported for La<sub>0.7</sub>Sr<sub>0.3</sub>MnO<sub>3</sub>.
We conclude that is not only the ferromagnetic interaction that is causing the increase in magnetization in La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub>. In order to understand this behavior we have to include other specific details of the crystalline and magnetic structure of the charge ordered material. These intriguing results could be a consequence of the competition between ferromagnetic double exchange and antiferromagnetic superexchange coupling. However, further studies in other charge ordering compounds, as well as simultaneous measurements of magnetization, resistivity, etc. are desirable to elucidate this unusual behavior.
We thank FAPESP, CAPES, CNPq and PRONEX for financial support. We also thank professor P. Nozar for a helpful discussion.
Figure 1. Magnetization per Mn ion versus temperature measured with an applied magnetic field of 5 T (1.2 mT in the inset) for a La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> sample. Magnetization is given in Bohr magnetons per manganese ion. Arrows and numbers show the direction of temperature sweep.
Figure 2. Normalized magnetic relaxation measurements after applying and removing an applied magnetic field of 5 T. Time is shown in logarithmic scale. The large arrow indicates the direction of increasing temperatures and relaxation rates (slope less negative) between 150 K and 195 K. Also shown is the curve for 145 K that presents a higher relaxation rate in comparison with 150 K. The inset reproduces details of the curve at 195 K with the corresponding error bars.
Figure 3. Normalized magnetic relaxation measurements after applying and removing a magnetic field of 5 T. Time is shown in logarithmic scale. The large arrow indicates the direction of increasing temperatures between 195 K and 240 K. Also shown is the curve for 245 K that presents a smaller increase of the normalized magnetization. The inset reproduces details of the curve at 200 K with the corresponding error bars (same dimension of circles), showing an unusual increase in the magnetization.
Figure 4. Magnetic relaxation curves at 210 K after repeating the standard procedure but reducing the field from 5 T to H=0 mT (squares), H=-1.2 mT (circles) and H=-10 mT (triangles in the inset). The initial magnetization of the sample decreases with increasing negative magnetic field and all curves show an increase in magnetization with time. Error bars are of the same dimension of symbols for all curves.
Figure 5. Normalized magnetic relaxation measurements at 210 K after applying a 5 T magnetic field, keeping it for a waiting time t<sub>w</sub>= 50, 500 and 5000 s, and removing it. The increase in normalized magnetization is higher the longer the magnetic field remains applied. |
warning/0003/hep-th0003065.html | ar5iv | text | # Untitled Document
Field Theory as a Matrix Model
Vladimir Kazakov kazakov@physique.ens.fr
<sup>1</sup> Laboratoire de Physique Thรฉorique de lโEcole Normale Supรฉrieure Unitรฉ Mixte du Centre National de la Recherche Scientifique et de lโEcole Normale Supรฉrieure.
75231 Paris CEDEX, France
A new formulation of four dimensional quantum field theories, such as scalar field theory, is proposed as a large $`n`$ limit of a special $`n\times n`$ matrix model. Our reduction scheme works beyond planar approximation and applies for QFT with finite number of fields. It uses quenched coordinates instead of quenched momenta of the old Eguchi-Kawai reduction known to yield correctly only the planar sector of quantum field theory. Fermions can be also included.
LPTENS-00/10
February, 2000
1. Introduction
To reproduce the physical space-space time out of some more fundamental variables, rather than introduce it explicitly, has been always a tempting idea in the quantum field theory. While the string theory sets up a fruitful framework in which the space-time is dynamically created out of the fluctuating coordinates of the strings, potentially not less fruitful may be the attempts to encode the space-time, together with the quantum fields themselves, into the dynamics of large fluctuating matrices. A quantum field theory or a string theory should be described in this case in terms of specific matrix integrals containing finite amount of matrices of infinite size. The earliest proposal of this kind belongs to T. Eguchi and H. Kawai , followed by a few important precisions and modifications ,,, . In these works the QFT with $`N\times N`$ matrix valued fields can be reduced in the large N (planar, or โt Hooft) limit to finite dimensional matrix integrals in the same limit.
Another successful enterprise of this kind was a formulation and solution of non-critical string theories (associated with the two dimensional quantum gravity in the presence of some matter fields) in terms of $`U(N)`$ invariant matrix integrals , .
More recently, one of the most fruitful ideas of this kind in the superstring theory was the proposal of E. Witten generalizing an old idea of Chan-Paton to describe the low energy physics of branes by various reductions of super Yang-Mills theory. The diagonal components of the effective SYM vector potentials play the role of space-time coordinates of branes there.
The concept of the physical space-time built out of discrete degrees of freedom has become especially inspiring due to fundamental questions in quantum gravity, such as microscopic explanation of the thermodynamics of black holes.
Our discussion in this paper will be confined to the matrix model formulation of quantum field theory. To set up the problem let us recall that the old Eguchi-Kawai (EK) reduction <sup>1</sup> Using this abbreviation we remember of course about the important modification of the original, not quite working, reduction scheme of Eguchi and Kawai related to the quenching of momenta and twisting reproduces correctly only the planar sector of a matrix field theory, whereas the non-planar corrections ($`1/N`$ expansion) were never incorporated into this scheme, let alone a nonperturbative formulation of a QFT with finite number of field components (finite N) in terms of some matrix model of a matrix or matrices of infinite size. The reason for this difficulty is mostly due to the lack of reduced momenta running around nontrivial cycles of non-planar graphs in the EK reduction scheme.
The aim of the present paper is to propose a new formulation of a finite component scalar field theory (finite $`N`$) in terms of an $`n\times n`$ matrix integral in the limit $`n\mathrm{}`$. One can say that it incorporates all orders of the $`1/N`$ expansion and is in principle a possible non-perturbative definition of the original scalar field theory. For example, the usual scalar $`\varphi ^4`$ theory can be formulated as a one matrix integral in external matrix sources. We will also show how to incorporate fermions into this scheme. Unfortunately, we did not find so far any natural way to formulate the four dimensional QCD in this way.
Our construction is in some sense T-dual to the old EK scheme: we use the diagonal matrix sources of quenched coordinates instead quenched momenta of the EK scheme. As a consequence of it the original scalar field โlivesโ on the graphs dual to feynman graphs of our matrix model. To control the parameters of these graphs (say, to make them exactly $`\varphi ^4`$ graphs) we apply the methods worked out in and for the so called model of dually weighted graphs (DWG).
We will be able to generalize our method to fermions and to their yukawa interactions with the scalars, but for the moment we donโt know a natural way to introduce gauge symmetry into our approach. So to formulate QCD is an interesting challenge in our framework.
2. Difficulty with $`1/N`$ corrections in the EK reduction scheme
Let us remind the essence of the old (and unsolved, to our knowledge) problem of $`1/N`$ corrections to the reduced version of planar field theory. We will mostly discuss a matrix version of scalar field theory in the euclidean four dimensional space described by the action
$$S=Nd^4x\mathrm{tr}\left((_\mu \varphi )^2+V(\varphi )\right)$$
where $`\varphi _{nm}(x)`$ is an $`N\times N`$ hermitian matrix field and $`V(\varphi )=_{k2}\frac{1}{k}t_k\varphi ^k`$ is a scalar potential.
The EK reduction (in the most natural, Parisi formulation ) goes as follows: take the following particular dependence of $`\varphi `$ on the coordinates:
$$\varphi _{mn}(x)=e^{i(p_mx)}\varphi _{mn}e^{i(p_nx)},\mathrm{๐๐}\mathrm{๐๐๐๐๐๐๐๐๐}\mathrm{๐๐๐๐}๐,๐$$
The action then takes the form
$$S=๐ฑ\mathrm{tr}\left([p_\mu ,\varphi ]^2+V(\varphi )\right)$$
where $`๐ฑ`$ is the 4D volume of the physical space, $`p_\mu =diag(p_\mu ^{(1)},\mathrm{},p_\mu ^{(N)})`$ are D diagonal matrices of quenched momenta scattered uniformly in a large 4D โmomentum boxโ of a size $`\mathrm{\Lambda }^4`$, where $`\mathrm{\Lambda }`$ is UV cutoff, and the original functional integral over scalar field is replaced by a single matrix integral over $`x`$-independent $`\varphi _{mn}`$. The planar sectors (leading large N approximation) of the two models, the original matrix scalar field theory in 4D and the reduced 0D one matrix model, coincide. It is immediately clear from the double line representation of planar graphs in the reduced theory (fig.1): each face of such a graph is associated with a closed index loop and with the momentum variable $`p_\mu ^{(i)}`$carrying the same index $`i`$; each double line propagator $`D_{jj^{}}^{ii^{}}=\frac{1}{N}\frac{\delta _{ii^{}}\delta _{jj^{}}}{(p_ip_j)^2}`$ depends on the difference of the momenta of adjacent faces.
Fig.1: A fragment of a planar diagram in the double line notation and the propagator depending on the quenched momenta $`p_k,p_j,k=1,\mathrm{},N`$ in Eguchi-Kawai reduction scheme.
The planar part of the free energy of the reduced model $`F_{plan}=\mathrm{๐๐๐}_N\mathrm{}\frac{1}{N^2}\mathrm{log}Z_N`$ ( $`Z_N`$ is the partition function) can be schematically written in the following way:
$$F_{plan}=๐ฑ\underset{G}{}\underset{v}{}t_{k_v}\left(\underset{\stackrel{~}{v}}{}\frac{1}{N}\underset{i_{\stackrel{~}{v}}}{}\right)\underset{<\stackrel{~}{v}\stackrel{~}{v}^{}>}{}\frac{1}{(p_{i_{\stackrel{~}{v}}}p_{j_{\stackrel{~}{v}^{}}})^2}$$
where $`_G`$ goes over planar feynman graphs $`G`$ of the scalar field theory, $`v,\stackrel{~}{v}`$ we label respectively original and dual vertices of a graph $`G`$, $`i_{\stackrel{~}{v}}1,\mathrm{},N`$ is the index associated with the dual vertex $`\stackrel{~}{v}`$ and $`_{<\stackrel{~}{v}\stackrel{~}{v}^{}>}`$ goes over the edges $`<\stackrel{~}{v}\stackrel{~}{v}^{}>`$ of the graph $`\stackrel{~}{G}`$ dual to $`G`$.
In the large N limit only planar graphs survive in both models and the sums over indices reproduce the integrals over 4D momenta: $`\frac{1}{N}_{i=1}^N\mathrm{}_{|p|<\mathrm{\Lambda }}d^4p\mathrm{}`$. Hence the planar sectors of the original matrix scalar field theory and of its reduced version are equivalent. This equivalence extends to any one point $`U(N)`$ invariant physical quantities of the type $`O_k=<\frac{tr}{N}\varphi ^k(x)>`$ but is known to fail for the multi-point correlators since $`<\frac{tr}{N}\varphi ^k(x)\frac{tr}{N}\varphi ^l(y)>=<\frac{tr}{N}\varphi ^k(x)><\frac{\mathrm{tr}}{N}\varphi ^l(y)>`$ plus $`O(1/N^2)`$ corrections, different in two models. The EK reduction fails to describe correctly the higher $`1/N`$ (non-planar) corrections to the original scalar field theory.
The reason for this failure is well known: we cannot represent all momenta running through the propagators as differences of momenta of the adjacent loops (faces) on the graphs of a non-spherical topology. If we did so (and it is precisely the case of the topological expansion in the EK reduced model) the momenta running along topologically nontrivial cycles of a non-planar feynman graph would be zero. To see it one takes any nontrivial closed path on a dual graph (connecting dual vertices or original faces) and calculates the total momentum running through the propagators crossed by this path as $`(p_ip_j)+(p_jp_k)+\mathrm{}+(p_lp_i)0`$. For example, for the torus topology we have two momenta of the original 4D theory missing in the reduced version: they flow through two nontrivial cycles of the torus.
The reason for the failure is simple but the remedy is not easy to find, at least in case of the EK reduction involving reduced momenta.
In the next section we will show that the goal of construction of a matrix model describing a finite $`N`$ (including the most frequent $`N=1`$ case) scalar field theory can be achieved by introducing quenched coordinates instead of quenched momenta.
3. Matrix model formulation of finite $`N`$ scalar field theory
Now we will propose a one matrix model in external matrix fields which will be equivalent, at least perturbativly, graph by graph of any topology, to the original 4D finite $`N`$ matrix scalar field theory (2.1). We will show that the free energy and, with an appropriate definition, the physical quantities of the field theory (2.1) at finite $`N`$ coincide with those of the matrix integral over a hermitian $`n\times n`$ matrix $`\mathrm{\Phi }`$ in the limit $`n\mathrm{}`$:
$$Z=e^{N^2F}=d^{n^2}\mathrm{\Phi }e^S$$
with the action:
$$S=N\mathrm{Tr}_๐ฉ\left([X_\mu ,\mathrm{\Phi }]^2+\mathrm{ln}(I_๐ฉA\mathrm{\Phi })\right)$$
Here the matrix $`\mathrm{\Phi }`$ lives in the $`n=pq`$ dimensional vector space $`๐ฉ`$ which is a direct product $`๐ฉ=๐ซ\times ๐ฌ`$ of vector spaces of smaller dimensions $`p`$ and $`q`$, correspondingly. $`I_n`$ is the $`n\times n`$ unity matrix. Both dimensions $`p,q`$ go to infinity as $`n\mathrm{}`$; $`X_\mu `$ and $`A`$ are external (fixed) diagonal matrices of the form
$$X_\mu =\widehat{x}_\mu \times I_๐ฌ,$$
where $`\widehat{x}_\mu =diag(x_\mu ^{(1)},\mathrm{},x_\mu ^{(p)})`$,
$$A=I_๐ซ\times \widehat{a}$$
where $`\widehat{a}=diag(a_1,\mathrm{}a_q)`$.
So the matrices $`\widehat{x}_\mu `$ on the one hand and the matrix $`\widehat{a}`$ on the other hand live in orthogonal subspaces.
The matrices $`\widehat{x}_\mu `$ with $`\mu =1,2,3,4`$ will play the role of quenched coordinates: the points with coordinates $`x_\mu ^{(i)}`$ should be distributed uniformly in the physical space box of a size $`L^4`$, which is the size of our system. The ultraviolet cutoff is defined as $`\mathrm{\Lambda }\frac{p^{1/4}}{L}`$. Obviously the thermodynamic limit of (infinite volume) corresponds to $`p\mathrm{}`$ with $`\mathrm{\Lambda }`$ fixed.
The matrix $`\widehat{a}`$ will encode the information about the couplings of scalar potential
$$V(\varphi )=\underset{k2}{}\frac{1}{k}t_k\varphi ^k$$
in the form:
$$t_k=\frac{p}{N}\mathrm{tr}_Q\widehat{a}^k$$
where the trace $`\mathrm{tr}_๐ฌ`$ goes only with respect to the vector space $`๐ฌ`$. Here $`t_2m^2`$ corresponds to the mass squared.
The parametrization (3.1) of the couplings reminds the so called Miwa variables widely used in the theory of $`\tau `$-functions of the hierarchies of integrable differential equations. It is also used in the representation of characters of the group $`GL(N)`$ through Schur polynomials (see for the details).
Obviously the last formula can be in general true only in the limit $`q\mathrm{}`$. Note that $`N`$ is kept as a finite fixed parameter here.
ยฟFrom (3.1) and (3.1) the potential can be also written in the form
$$V(z)=\frac{p}{N}\underset{j=1}{\overset{q}{}}\mathrm{ln}(1a_jz)$$
In the limit $`q\mathrm{}`$ we can parameterize in principle any potential (including a polynomial one) by such a sum of logarithmic terms, but $`a_i`$โs need not be necessary real. They can be taken, say, in complex conjugate pairs (the potential becomes even in this case).
For instance we can choose $`a_i`$ in such a way that in the limit $`q\mathrm{}`$ they will reproduce any polynomial potential (3.1):
$$a_j=e^{i\theta _j}$$
where
$$\theta _j=\frac{2\pi }{q}j+\frac{2N}{pq}\underset{m1}{}\frac{t_m}{m}\mathrm{sin}\frac{2\pi jm}{q}$$
Indeed, for $`k>0`$ we have
$$\underset{j=1}{\overset{q}{}}e^{ik\theta _j}_{p,q\mathrm{}}\underset{j=1}{\overset{q}{}}e^{\frac{2i\pi }{q}jk}\left(1+ik\frac{2N}{pq}\underset{m1}{}\frac{t_m}{m}\mathrm{sin}\frac{2\pi jm}{q}\right)=\frac{N}{p}t_k.$$
The proof of the equivalence of the QFT (2.1) and the matrix model (3.1) is very simple. Let us consider any feynman graph of the theory (3.1) (dotted line on fig. 2) together with its dual graph (solid line on fig. 2). The matrix structure of the theory prescribes to use the double line notations for the propagators, so such graph has a fixed topology with the genus defined by the Euler formula. Each single line carries now a double index corresponding to the product of spaces $`๐ซ\times ๐ฌ`$.
Fig.2. Original and dual graphs and the dual propagator depending on the quenched coordinates $`x_k,x_j`$ in our one matrix integral representation of scalar QFT. We show the pairs of indices $`b,k`$ and $`c,j`$ belonging to $`๐ฌ,๐ซ`$ spaces, correspondingly, running around original faces, and hence placed at dual vertices.
The propagators of the original graph are given by:
$$\frac{1}{N}\frac{a_ba_c}{(x_ix_j)^2}\delta _{bb^{}}\delta _{cc^{}}\delta _{ii^{}}\delta _{jj^{}}.$$
Here we attributed the $`a_k`$ factors to the propagators rather than to the vertices which can be achieved by the change of the matrix variable: $`\mathrm{\Phi }A^{1/2}\mathrm{\Phi }A^{1/2}`$. It is easy to see that we obtain at each face of the original graph a weight $`\mathrm{tr}_๐ฌ\widehat{\alpha }^k`$ where k is the order of this face (number of edges). It happens in the same way as in the so called matrix model of dually weighted graphs (DWG) - a one matrix model with the action $`S=n\mathrm{tr}[\mathrm{\Phi }^2+W(A\mathrm{\Phi })]`$ studied in , and solved in : apart from original couplings coming from the potential $`W(z)`$ we also obtain in the DWG model the dual couplings $`\stackrel{~}{t}_k`$ weighting the faces of different orders k (or, which is the same, the vertices of the dual graph with the coordination number k).
The factors $`\frac{1}{(x_ix_j)^2}`$ can be now attributed to the dual propagators (crossing the original ones) and the indices $`i,j`$ are running around the faces adjacent to the original propagators or, which is the same, attributed to the corresponding dual vertices (see fig.2).
It is natural to formulate the result for the free energy in terms of feynman expansion with respect to the dual graphs $`\stackrel{~}{G}`$. It can be written in the following way:
$$F=\underset{\stackrel{~}{G}}{}N^{22g}\underset{\stackrel{~}{v}}{}\left(t_{k_{\stackrel{~}{v}}}\frac{1}{p}\underset{i_{\stackrel{~}{v}}=1}{\overset{p}{}}\right)\underset{<\stackrel{~}{v}^{}\stackrel{~}{v}^{\prime \prime }>}{}\left(x_{i_{\stackrel{~}{v}^{}}}x_{i_{\stackrel{~}{v}^{\prime \prime }}}\right)^2$$
where $`_{\stackrel{~}{G}}`$ goes over all dual graphs of the matrix model (3.1), $`g`$ is the genus of a graph $`\stackrel{~}{G}`$, $`_{\stackrel{~}{v}}`$ goes over all vertices $`\stackrel{~}{v}`$ of this graph with coordination numbers $`k`$, $`t_k=\frac{p}{N}_{j=1}^qa_j^k`$ are the couplings attached to these vertices and $`_{<\stackrel{~}{v}^{}\stackrel{~}{v}^{\prime \prime }>}`$ goes with respect to all edges of $`\stackrel{~}{G}`$ connecting the vertices $`\stackrel{~}{v}`$ and $`\stackrel{~}{v}^{}`$. At each dual vertex $`\stackrel{~}{v}`$ there is a sum taken with respect to the index $`i_{\stackrel{~}{v}}`$ corresponding to the subspace $`๐ซ`$.
Note also that due to the specific logarithmic form of the interactions in (3.1) $`\mathrm{Tr}\mathrm{ln}(IA\mathrm{\Phi })=_k\frac{1}{k}\mathrm{Tr}(A\mathrm{\Phi })^k`$ all faces of dual graphs appear weighted with the factor 1 (the factor $`\frac{1}{k}`$ compensates the cyclic symmetry of each dual face). So we see that (3.1) is given by the sum over connected graphs $`\stackrel{~}{G}`$ of all genera waited by $`N^{22g}`$, with unrestricted face order and with the vertices waited by couplings $`t_{k_{\stackrel{~}{v}}}=\frac{p}{N}\mathrm{tr}_๐ฌ\widehat{a}^{k_{\stackrel{~}{v}}}`$. Hence these are precisely the original graphs of the scalar matrix field theory (2.1) with the appropriate 4D massless propagators in the coordinate space. The mass $`m`$ is taken into account by the presence of the coupling $`t_2m^2`$ in the scalar potential.
It is left to add that the summations over the indices $`i_{\stackrel{~}{v}}=1,\mathrm{},p`$ can be substituted by the integrations in the large $`p`$ limit:
$$\frac{1}{p}\underset{i=1}{\overset{p}{}}\mathrm{}_p\mathrm{}d^4x\mathrm{}$$
In this way we can reproduce the correspondence between the matrix scalar 4D QFT (2.1) and the zero dimensional matrix integral (3.1), graph by graph of any topology. Hence they coincide, at least in any order of the perturbation theory.
Let us stress again that $`N`$ is a fixed parameter in our construction. It does not even need to be integer, although integer $`N`$ seems to be singled out by the fact that we can represent the corresponding determinant as an integral over an $`N`$-vector of complex $`n\times n`$ matrices $`M_l,l=1,\mathrm{},N`$:
$$\mathrm{exp}[N\mathrm{Tr}\mathrm{log}(IA\mathrm{\Phi })]=\underset{l=1}{\overset{N}{}}d^{2N^2}M_l\mathrm{exp}[\mathrm{Tr}M_l^+(IA\mathrm{\Phi })M_l]$$
In particular, for $`N=1`$ (3.1) is equivalent to the usual one component scalar field theory with the action $`S=d^4x\left[(_\mu \varphi )^2+V(\varphi )\right]`$.
To calculate the one point correlators we use the following correspondence between the averages in the scalar QFT and its matrix model (MM) formulation (we take $`N=1`$):
$$<d^4x\varphi ^k(x)>_{QFT}=\frac{1}{\mathrm{tr}_๐ฉA^k}<\mathrm{tr}_๐ฉ(A\mathrm{\Phi })^k>_{MM}$$
For the two point correlator we have the correspondence:
$$<\varphi (x)\varphi (y))>_{QFT}=\frac{1}{(\mathrm{tr}_๐ฌ\widehat{a})^2}<(\mathrm{tr}_๐ฉ(A\mathrm{\Phi })\left)_{ii}\right(\mathrm{tr}_๐ฉ(A\mathrm{\Phi }))_{jj}>_{MM}$$
where the traces are taken only with respect to the subspace $`๐ฌ`$ and the matrix indices $`ii`$ and $`jj`$ of the subspace $`๐ซ`$ are fixed and chosen in such a way that $`(x_ix_j)^2(xy)^2`$.
4. Reduction in the presence of fermions and yukawa interactions
Although we donโt know any natural way to build a reduction of Yang-Mills theory for finite $`N`$, the scalar field theory (2.1) is not the only interesting QFT which can be reduced to a matrix model in this way.
Let us consider as an example of application of our method the reduction of the QFT of massless Dirac fermions and massless bosons with yukawa interaction in 4D with the action:
$$S=d^4x\left\{(_\mu \varphi )^2+\overline{\psi }\gamma _\mu \psi \frac{\lambda }{3}\overline{\psi }\psi \varphi \right\}$$
We stress that here $`\varphi (x)`$ and $`\psi (x)`$ are usual bosonic and dirac fields: $`\varphi `$ has only one component and $`\psi `$ is a dirac spinor. The reduced version of this model is given in terms of matrix integral over 2 hermitian $`n\times n`$ matrices $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$ in the auxiliary linear space $`๐ฉ=๐ฌ\times ๐ซ`$
$$Z=e^F=d^{n^2}\mathrm{\Phi }d^{n^2}\mathrm{\Psi }e^S$$
with the action
$$S=\mathrm{Tr}_๐ฉ\left([X_\mu ,\mathrm{\Phi }]^2+[X_\mu ,[X_\nu ,\mathrm{\Psi }]]^2\mathrm{tr}_D\mathrm{ln}(I_๐ฉ\gamma _\mu \times [X_\mu ,\mathrm{\Psi }]A)+\mathrm{ln}(I_๐ฉA\mathrm{\Phi }A\mathrm{\Psi })\right)$$
Here $`\mathrm{tr}_D`$ is taken with respect to the dirac indices of $`\gamma _\mu `$ matrix which is in the direct product with all other matrices in the third term. The $`A`$ matrix is chosen in such a way that $`p_{j=1}^qa_j^k=\lambda \delta _{k,3}`$ for $`k>0`$ in the large $`q`$ limit. The last can be achieved for example by the following choice of $`\widehat{a}`$: $`a_j=\left(\frac{2\lambda }{pq}\right)^{1/3}e^{i\theta _j}`$ with $`\theta _j`$ given by the equation $`\frac{2\pi j}{q}=\theta +\frac{1}{3}\mathrm{sin}3\theta `$. To prove it it is enough to calculate the density of $`\theta `$โs in the large $`q`$ limit: $`\rho (\theta )=\frac{1}{q}\frac{j}{\theta }=\frac{1}{2\pi }(1+\frac{1}{3}\mathrm{cos}3\theta )`$ which gives $`\mathrm{tr}_๐ฌ\widehat{a}^k=_\pi ^\pi e^{ik\theta }\rho (\theta )=\lambda \delta _{k,3}`$ for $`k>0`$.
All other definitions are the same as in the previous section.
To verify the perturbative equivalence of the QFT (4.1) and the matrix integral (4.1)-(4.1) we compare again the feynman graphs of the former to the dual feynman graphs of the latter. Due to the choice of the matrix $`\widehat{a}`$ dual graphs of (4.1) will contain only triple interaction vertices and due to the $`\mathrm{log}`$ type potentials the order of dual faces will be unrestricted (see fig. 3).
Fig.3: Original and dual graphs of the QFT of scalars and fermions with yukawa interactions (in single line notations).
Note that feynman graphs of the QFT (4.1) have two types of faces: fermionic loops built out of Dirac propagators and the loops with the boundary built from interchanging dirac and bosonic propagators. The first type of loops will be generated by the third term in (4.1) and the second kind - by the last term in (4.1). The yukawa interaction vertices correspond to the loops of dual graphs of the matrix model (4.1). Note the $`()`$ sign in front of the third term which gives the fermionic statistics ($`()`$ sign for each fermionic loop). Each fermionic loop of $`k`$ vertices is equipped, as it should be, by the factor $`\mathrm{tr}_D(\gamma _{\mu _1}\mathrm{}\gamma _{\mu _k})`$. Finally, the massless 4D fermionic propagator $`\frac{\gamma _\mu (x^\mu y^\nu )}{|xy|^4}`$ will appear as a function of difference of quenched coordinates $`\frac{\gamma _\mu (x_i^\mu x_i^\nu )}{|x_ix_j|^4}`$.
Hence in the same limit $`p,q\mathrm{}`$ as for the scalar QFT of the previous section we reproduce graph by graph (with the weights independent on the topology of graphs) the free energy of the QFT (4.1) out of the matrix integral (4.1).
5. Conclusions and comments
We proposed a reduction of 4D quantum field theories with a finite number of fields (finite $`N`$ in case of matrix fields) to a matrix integral over infinite matrices. The physical four dimansional coordinate space is encoded into the components of D=4 auxiliary diagonal matrices of quenched coordinates. The reduction is different from the old Eguchi-Kawai reduction with quenched momenta: it reproduces correctly not only the planar approximation but also, at least perturbatively graph by graph, all non planar corrections. The coordinate space is introduced in our reduction scheme in a way which reminds the description of coordinates of D-branes in . The large $`n`$ limit that we use is different from the usual โtHooft limit and rather similar to the one adopted in for the matrix model of M-theory.
We showed that fermions can be also naturally introduced into this reduction scheme.
A few comments are in order:
1. The QFTโs in dimesions different from D=4 can be also reduced in this way but since the scalar propagator is different from $`1/(xy)^2`$ the corresponding matrix representation of it looks โnonlocalโ (i.e., different from $`\mathrm{Tr}[X_\mu ,\mathrm{\Phi }]^2`$). One can use the formula $`\mathrm{Tr}[X_{\mu _1},[\mathrm{},[X_{\mu _k},\mathrm{\Phi }]]\mathrm{}]^2=_{i,j}(x_ix_j)^{2k}|\mathrm{\Phi }_{ij}|^2`$ in the action of the reduced matrix model to supply the dual graphs by any propagator $`D(x_ix_j)`$.
2. Although our new matrix formulation of some QFTโs hardly helps for solving them analytically it may provide a new numerical approach to their study: our matrix model does not deal with any 4D lattice and the approach to the thermodynamical limit might be much faster then in the conventional Monte Carlo algorithms based on the lattices. One could also envisage some real space renormalization schemes where the renormalization flow would be considered with respect to the size of matrices.
3. The authors of the papers , propose a formulation of the non-comutative QFT using reduction to the matrix models with quenched momentum matrices $`P_\mu `$ obeying the Heisenberg commutation relations: $`[P_\mu ,P_\nu ]=iB_{\mu \nu }`$ where $`B_{\mu \nu }`$ is an antisymmetric $`D\times D`$ matrix of C-numbers. For our model, a natural non-comutative generalization would occur if we take the same commutation relations for the coordinate matrices $`X_\mu `$: $`[X_\mu ,X_\nu ]=iC_{\mu \nu }`$. Since the coordinate and momentum matrices are connected in by the relation $`P_\mu =B_{\mu \nu }X_\nu `$ it is natural to expect that we get the same non-comutative scalar field theory if we take the matrix $`C=B^1`$. We havenโt yet a proof of this statement.
Let us also mention the most obvious problems and questions concerning our matrix reduction of QFTโs:
1. We did not manage to find a natural matrix reduction of QCD with 3 colors. In principle we could try to do it in a particular gauge by systematically reproducing all feynman diagrams by the method we presented in this paper. But it would be much more instructive to find some general principle for such reduction arising explicitely from the gauge invariance. For the moment such a principle is missing.
2. It would be interesting to find a modification of our approach similar to the twisted EK model proposed in . In the momentum space appears as a classical vacuum of a large N lattice QFT with modified couplings. We can imagine a similar twisting of reduced matrix model with the classical vacuum solution generating the coordinate space.
3. A related but more ambitious question: can we build realistic models of fundamental interactions as some reduced matrix models where the physical space would emerge due to some symmetry breaking procedure? In other word, can our World be described by a Matrix Model? We know some of the attempts of this kind in the superstring theory and M-theory , . Our construction might be useful to approach this problem from a different direction.
6. Acknowledgments
The discussions with Hong Liu and Ivan Kostov were very useful for this project.
I would like to thank the Physics Department of the Universidad Catolica in Santiago (Chile) where this work was started, and especially Jorge Alfaro for the kind hospitality.
I am particularly grateful to the Weizmann Institute for Science in Israel where most of this work was done during my stay there on the Landau-Weizmann program and especially Shimon Levit and Adam Schwimmer for their kind hospitality and fruitful interactions.
This research is supported in part by European TMR contract ERBFMRXCT960012 and EC Contract FMRX-CT96-0012.
References
relax T. Eguchi and H. Kawai, Phys. Rev. Lett. 48 (1982) 1063. relax G. Bhanot, U. Heller and H. Neuberger, Phys. Lett.113B (1982) 47. relax G. Parisi, Phys. Lett.113B (1982) 463 relax D. Gross and Y. Kitazawa, Nucl. Phys. B206 (1982) 440. relax A. Gonsalez-Arroyo and M. Okawa, Phys. Lett.B120 (1983) 174, Phys.Rev. D27 (1983) 2397. relax F. David, Nucl. Phys. B257 (1985) 45. relax V. A. Kazakov, Phys. Lett.150B (1985) 282. relax E. Witten, Nucl. Phys. B460 (1995) 335, hep-th/9510135. relax C. Itzykson and P. DiFranchesco, Ann. Inst. Henri Poincarรฉ, vol.59, no. 2 (1993) 117. relax V. A. Kazakov, M. Staudacher and T. Wynter, Comm. Math. Phys. 177 (1996) 451, hep-th/9502132; Comm. Math. Phys. 179 (1996) 235, hep-th/9506174; Nucl. Phys. B440 (1995) 407, hep-th/9601069. relax S. R. Das, A. Dhar, A. Sentgupta and S. R. Wadia, Mod. Phys. Lett.A5 (1990) 41. relax Banks, W. Fischler, S. Shenker and L. Susskind, Phys.Rev. D55 (1997) 5112, hep-th/9610043. relax N. Ishibashi, S. Iso, H. Kawai and Y. Kitazawa, hep-th/9910004, hep-th/0001027. relax J. Ambjorn, Yu. Makeenko, J. Nishimura and R. J. Szabo, hep-th/9911041. relax N. Ishibashi, H. Kawai, Y. Kitazawa and A. Tsuchiya, Nucl. Phys. B498 (1997) 467, hep-th/9612115. |
warning/0003/math-ph0003044.html | ar5iv | text | # Contents
## 1 Introduction
One of the basic principles of modern theoretical physics is the principle of local gauge invariance. Its application to the theory of particle interactions gave rise to the standard model, which proved to be a success from both theoretical and phenomenological points of view. The most impressive results of the model were obtained within the perturbation theory, which works well for high energy processes. On the other hand, the low energy hadron physics, in particular, the quark confinement, turns out to be dominated by nonperturbative effects, for which there is no rigorous theoretical explanation yet.
The application of geometrical methods to non-abelian gauge theories revealed their rich geometrical and topological properties. In particular, it showed that the configuration space of such theories, which is the space of gauge group orbits in the space of connections, may have a highly nontrivial structure. In general, the orbit space possesses not only orbits of the so called principal type, but also orbits of other types, which may give rise to singularities of the configuration space. This stratified structure of the gauge orbit space is believed to be of importance for both classical and quantum properties of non-abelian gauge theories in nonperturbative approach, and it has been intensively studied in recent years. Let us discuss some aspects indicating its physical relevance.
First, studying the geometry and topology of the generic (principal) stratum, one gets a deeper understanding of the Gribov-ambiguity and of anomalies in terms of index theorems. In particular, one gets anomalies of purely topological type, which can not be seen by perturbative quantum field theory. These are well-known results from the eighties. Moreover, there are partial results and conjectures concerning the relevance of nongeneric strata. First of all, nongeneric gauge orbits affect the classical motion on the orbit space due to boundary conditions and, in this way, may produce nontrivial contributions to the path integral. They may also lead to localization of certain quantum states, as it was suggested by finite-dimensional examples . Further, the gauge field configurations belonging to nongeneric orbits can possess a magnetic charge, i.e. they can be considered as a kind of magnetic monopole configurations, which are responsible for quark confinement. This picture was found in 3-dimensional gauge systems , and it is conjectured that it can hold for 4-dimensional theories as well . Finally, it was suggested in that non-generic strata may lead to additional anomalies.
Most of the problems mentioned here are still awaiting a systematic investigation. In a series of papers we are going to make a new step in this direction. We give a complete solution to the problem of determining the strata that are present in the gauge orbit space for $`\mathrm{SU}n`$ gauge theories in compact Euclidean space-time of dimension $`d=2,3,4`$. Our analysis is based on the results of a paper by Kondracki and Rogulski , where it was shown that the gauge orbit space is a stratified topological space in the ordinary sense, cf. and references therein. Moreover, these authors found an interesting relation between orbit types and certain bundle reductions, which we are going to use. We also refer to for the discussion of a very simple, but instructive special example (orbit types of $`\mathrm{SU2}`$-gauge theory on $`\mathrm{S}^4`$).
The paper is organized as follows. In Section 2 we introduce the basic notions related to the action of the group of gauge transformations on the space of connections, state the definitions of stabilizer and orbit type and recall basic results concerning the stratification structure of the gauge orbit space. In Section 3 we introduce the notions of Howe subgroup and holonomy-induced Howe subbundle and establish the relation between orbit types and holonomy-induced Howe subbundles. Section 4 is devoted to the study of the Howe subgroups of $`\mathrm{SU}n`$. In Section 5 we give a classification of the Howe subbundles of $`\mathrm{SU}n`$-bundles for space-time dimension $`d4`$. In Section 6 we prove that any Howe subbundle of $`\mathrm{SU}n`$-bundles is holonomy-induced. In Section 7 we implement the equivalence relation of Howe subbundles due to the action of $`\mathrm{SU}n`$. As an example, in Section 8 we determine the orbit types for gauge group $`\mathrm{SU2}`$. Finally, in Section 9 we discuss an application to Chern-Simons theory in $`2+1`$ dimensions.
In two subsequent papers we shall investigate the natural partial ordering on the set of orbit types and the structure of another, coarser stratification, see , obtained by first factorizing with respect to the so called pointed gauge group and then by the structure group.
## 2 Gauge Orbit Types and Stratification
We consider a fixed topological sector of a gauge theory with gauge group $`G`$ on a Riemannian manifold $`M`$. In the geometrical setting it means that we are given a smooth right principal fibre bundle $`P`$ with structure group $`G`$ over $`M`$. $`G`$ is assumed to be a compact connected Lie group and $`M`$ is assumed to be compact, connected, and orientable.
Denote the sets of connection forms and gauge transformations of $`P`$ of Sobolev class $`W^k`$ (locally square integrable up to the $`k`$-th derivative) by $`๐^k`$ and $`๐ข^k`$, respectively. Here $`k`$ is a nonnegative integer. For generalities on Sobolev spaces of cross sections in fibre bundles, see . Provided $`2k>dimM`$, $`๐^k`$ is an affine Hilbert space and $`๐ข^{k+1}`$ is a Hilbert Lie group acting smoothly from the right on $`๐^k`$ . We shall even assume $`2k>dimM+2`$. Then, by the Sobolev Embedding Theorem, connection forms are of class $`C^1`$ and, therefore, have continuous curvature. If we view elements of $`๐ข^{k+1}`$ as $`G`$-space morphisms $`PG`$, the action of $`g๐ข^{k+1}`$ on $`A๐^k`$ is given by
$$A^{(g)}=g^1Ag+g^1\mathrm{d}g.$$
(1)
Let $`^k`$ denote the quotient topological space $`๐^k/๐ข^{k+1}`$. This space represents the configuration space of our gauge theory.
For this to make sense, $`^k`$ should not depend essentially on the purely technical parameter $`k`$. Indeed, let $`k^{}>k`$. Then one has natural embeddings $`๐ข^{k^{}+1}๐ข^{k+1}`$ and $`๐^k^{}๐^k`$. As a consequence of the first, the latter projects to a map $`\phi :^k^{}^k`$. Since the image of $`๐^k^{}`$ in $`๐^k`$ is dense, so is $`\phi \left(^k^{}\right)`$ in $`^k`$. To see that $`\phi `$ is injective, let $`A_1,A_2๐^k^{}`$ and $`g๐ข^{k+1}`$ such that $`\phi (A_2)=\phi (A_1)^{(g)}`$. Then (1) implies
$$\mathrm{d}g=g\phi (A_2)\phi (A_1)g.$$
(2)
Due to $`2k^{}>2k>dimM`$, by the multiplication rule for Sobolev functions, the rhs. of (2) is of class $`W^{k+1}`$. Then $`g`$ is of class $`W^{k+2}`$. This can be iterated until the rhs. is of class $`W^k^{}`$. Hence, $`g๐ข^{k^{}+1}`$, so that $`A_1`$ and $`A_2`$ are representatives of the same element of $`^k^{}`$. This shows that $`^k^{}`$ can be identified with a dense subset of $`^k`$. Another question is whether the stratification structure of $`^k`$, which will be discussed in a moment, depends on $`k`$. Fortunately, the answer to this question is negative, see Theorem 3.3.
In general, the orbit space of a smooth Lie group action does not admit a smooth manifold structure w.r.t. which the projection is smooth. The best one can expect is that it admits a stratification. For the notion of stratification of a topological space, see or \[23, ยง4.4\]. For the gauge orbit space $`^k`$, a stratification was constructed in , using a method which is known from compact Lie group actions on completely regular spaces . In order to explain this, let us recall the notions of stabilizer and orbit type. The stabilizer, or isotropy subgroup, of $`A๐^k`$ is the subgroup
$$๐ข_A^{k+1}=\{g๐ข^{k+1}|A^{(g)}=A\}$$
of $`๐ข^{k+1}`$. It has the following transformation property: For any $`A๐^k`$ and $`g๐ข^{k+1}`$,
$$๐ข_{A^{(g)}}^{k+1}=g^1๐ข_A^{k+1}g.$$
Thus, there exists a natural map, called type map, assigning to each element of $`^k`$ the conjugacy class in $`๐ข^{k+1}`$ made up by the stabilizers of its representatives in $`๐^k`$. Let $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ denote the image of this map. The elements of $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ are called orbit types. The set $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ carries a natural partial ordering: $`\tau \tau ^{}`$ iff there are representatives $`๐ข_A^{k+1}`$ of $`\tau `$ and $`๐ข_A^{}^{k+1}`$ of $`\tau ^{}`$ such that $`๐ข_A^{k+1}๐ข_A^{}^{k+1}`$. Note that this definition is consistent with but not with and several other authors who define it just inversely.
As was shown in , the subsets $`_\tau ^k^k`$, consisting of gauge orbits of type $`\tau `$, can be equipped with a smooth Hilbert manifold structure and the family
$$\left\{_\tau ^k\right|\tau \mathrm{OT}(๐^k,๐ข^{k+1})\}$$
(3)
is a stratification of $`^k`$. Accordingly, the manifolds $`_\tau ^k`$ are called strata. In particular,
$$^k=\underset{\tau \mathrm{OT}(๐^k,๐ข^{k+1})}{}_\tau ^k,$$
where for any $`\tau \mathrm{OT}(๐^k,๐ข^{k+})`$ there holds
$`_\tau ^k`$ is open and dense in $`_{\tau ^{}\tau }_\tau ^{}^k`$. (4)
Similarly to the case of compact Lie groups, there exists a principal orbit type $`\tau _0`$ obeying $`\tau _0\tau `$ for all $`\tau \mathrm{OT}(๐^k,๐ข^{k+1})`$. Due to (4), the corresponding stratum $`_{\tau _0}^k`$ is open and dense in $`^k`$. For this reason it is called the generic stratum.
The above considerations show that the set $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ together with its natural partial ordering carries the information about which strata occur and how they are patched together.
To conclude, let us remark that instead of using Sobolev techniques one can also stick to smooth connection forms and gauge transformations. Then one obtains essentially analogous results about the stratification of the corresponding gauge orbit space where, roughly speaking, one has to replace โHilbert manifoldโ and โHilbert Lie groupโ by โtame Frรฉchet manifoldโ and โtame Frรฉchet Lie groupโ, see .
## 3 Correspondence between Orbit Types and Bundle Reductions
In this section, let $`p_0P`$ be fixed. For $`A๐^k`$, let $`H_A`$ and $`P_A`$ denote the holonomy group and holonomy subbundle, respectively, of $`A`$ based at $`p_0`$. We assume $`2k>dimM+2`$. Then, by the Sobolev Embedding Theorem, $`A`$ is of class $`C^1`$ so that $`P_A`$ is a subbundle of $`P`$ of class $`C^2`$. In particular, it is a subbundle of class $`C^0`$.
For any $`g๐ข^{k+1}`$, let $`\vartheta _g`$ denote the associated vertical automorphism of $`P`$. Then
$$\vartheta _g(p)=pg(p)pP.$$
(5)
Let $`\mathrm{C}_G(H)`$ denote the centralizer of $`HG`$ in $`G`$. We abbreviate $`\mathrm{C}_G^2(H)=\mathrm{C}_G\left(\mathrm{C}_G(H)\right)`$.
Let $`A๐^k`$ and $`g๐ข^{k+1}`$. Since the elements of $`๐ข_A^{k+1}`$ map $`A`$-horizontal paths in $`P`$ to $`A`$-horizontal paths they are constant on $`P_A`$. Conversely, any gauge transformation which is constant on $`P_A`$ leaves $`A`$ invariant. Thus, for any $`g๐ข^{k+1}`$ one has
$$g๐ข_A^{k+1}g|_{P_A}\text{ is constant.}$$
(6)
This suggests the idea to characterize orbit types by certain classes of subbundles of $`P`$. For any subgroup $`S๐ข^{k+1}`$ define a subset $`\mathrm{\Theta }(S)P`$ by
$$\mathrm{\Theta }(S)=\left\{pP\right|g(p)=g(p_0)gS\}.$$
(7)
###### Lemma 3.1
(a) For any $`A๐^k`$, $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)=P_A\mathrm{C}_G^2\left(H_A\right)`$.
(b) Let $`A,A^{}๐^k`$. Then $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)=\mathrm{\Theta }\left(๐ข_A^{}^{k+1}\right)`$ implies $`๐ข_A^{k+1}=๐ข_A^{}^{k+1}`$.
(c) Let $`g๐ข^{k+1}`$. For any subgroup $`S๐ข^{k+1}`$, $`\mathrm{\Theta }\left(gSg^1\right)=\vartheta _g\left(\mathrm{\Theta }(S)\right)g(p_0)^1`$.
Remark: In , $`P_A\mathrm{C}_G^2(H_A)`$ is called the evolution bundle generated by $`A`$. Thus, in this terminology, (a) states that $`\mathrm{\Theta }`$ maps the stabilizer of a connection to its evolution bundle.
Proof:
(a) Let $`A๐^k`$. Recall that $`P_A`$ has structure group $`H_A`$. Hence, in view of (6), the equivariance property of gauge transformations implies
$$\left\{g(p_0)\right|g๐ข_A^{k+1}\}=\mathrm{C}_G(H_A).$$
(8)
Thus, by equivariance again,
$$g๐ข_A^{k+1}g|_{P_A\mathrm{C}_G^2(H_A)}\text{ is constant.}$$
(9)
This shows $`P_A\mathrm{C}_G^2(H_A)\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$. Conversely, let $`pP`$ such that $`g(p)=g(p_0)`$ for all $`g๐ข_A^{k+1}`$. There exists $`aG`$ such that $`pa^1P_A`$. Then (6) yields
$$g(p_0)=g\left(pa^1\right)=ag(p)a^1=ag(p_0)a^1,$$
for all $`g๐ข_A^{k+1}`$. Due to (8), then $`a\mathrm{C}_G^2(H_A)`$. Hence, $`p=(pa^1)aP_A\mathrm{C}_G^2(H_A)`$.
(b) Let $`A,A^{}๐^k`$ be given. For any $`g๐ข^{k+1}`$, we have
$$g๐ข_A^{k+1}g|_{\mathrm{\Theta }\left(๐ข_A^{k+1}\right)}\text{ is constant.}$$
Here implication from left to right is due to (9) and assertion (a), the inverse implication follows from $`P_A\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ and (6). Since a similar characterization holds for $`๐ข_A^{}^{k+1}`$, the assertion follows.
(c) Let $`pP`$, $`hS`$. Using (5) we compute
$$g(p_0)^1g(p)h(p)g(p)^1g(p_0)=h\left(\vartheta _{g^1}(p)g(p_0)\right).$$
This allows us to write down the following chain of equivalences:
$$\begin{array}{ccc}\hfill p\mathrm{\Theta }\left(gSg^1\right)& & g(p)h(p)g(p)^1=g(p_0)h(p_0)g(p_0)^1\hfill \\ & & g(p_0)^1g(p)h(p)g(p)^1g(p_0)=h(p_0)\hfill \\ & & h\left(\vartheta _{g^1}(p)g(p_0)\right)=h(p_0)\hfill \\ & & \vartheta _{g^1}(p)g(p_0)\mathrm{\Theta }(S).\hfill \end{array}$$
This proves assertion (c).
###### Definition 3.2
Let $`G`$ be a Lie group and let $`P`$ be a principal $`G`$-bundle. A subgroup $`HG`$ is called Howe subgroup iff there exists a subset $`KG`$ such that $`H=\mathrm{C}_G(K)`$. A reduction of $`P`$ to a Howe subgroup of $`G`$ will be called Howe subbundle. A subbundle $`QP`$ will be called holonomy-induced of class $`C^r`$ iff it contains a connected subbundle $`\stackrel{~}{Q}Q`$ of class $`C^r`$ with structure group $`\stackrel{~}{H}`$ such that
$$Q=\stackrel{~}{Q}\mathrm{C}_G^2\left(\stackrel{~}{H}\right).$$
(10)
We remark that the notion of Howe subgroup is common in the literature, cf. . The notions of Howe subbundle and holonomy-induced subbundle are, to our knowledge, new. Moreover, we note that $`HG`$ is a Howe subgroup iff $`\mathrm{C}_G^2(H)=H`$. In particular, the subset $`KG`$ in the definition can be chosen to be $`\mathrm{C}_G(H)`$.
The set of holonomy-induced Howe subbundles of $`P`$ of class $`C^0`$ is acted upon in a natural way by the group of continuous gauge transformations $`๐ข^{C^0}`$ and by the structure group $`G`$. Let $`\text{Howe}_{}(P)`$ denote the corresponding set of conjugacy classes. We note that the actions of $`๐ข^{C^0}`$ and $`G`$ commute. Moreover, two subbundles of class $`C^0`$ of $`P`$ are conjugate under the action of $`๐ข^{C^0}`$ iff they are isomorphic. $`\text{Howe}_{}(P)`$ carries a natural partial ordering, which is defined similarly to that of orbit types: $`\eta \eta ^{}`$ iff there are representatives $`Q`$ of $`\eta `$ and $`Q^{}`$ of $`\eta ^{}`$ such that $`QQ^{}`$.
###### Theorem 3.3
Let $`M`$ be compact, $`dimM2`$. Then the assignment $`\mathrm{\Theta }`$ induces, by passing to quotients, an order-preserving bijection from $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ onto $`\text{Howe}_{}(P)`$.
Proof: Let $`\tau \mathrm{OT}(๐^k,๐ข^{k+1})`$ and choose a representative $`S๐ข^{k+1}`$. There exists $`A๐^k`$ such that $`S=๐ข_A^{k+1}`$. According to Lemma 3.1(a), $`\mathrm{\Theta }(S)`$ can be obtained by extending the subbundle $`P_AP`$ to the structure group $`\mathrm{C}_G^2(H_A)`$. Since $`P_A`$ is of class $`C^0`$, so is $`\mathrm{\Theta }(S)`$. Since $`P_A`$ is connected, $`\mathrm{\Theta }(S)`$ is holonomy-induced of class $`C^0`$. Moreover, it is obviously Howe. According to Lemma 3.1(c), when passing to quotients the class of $`\mathrm{\Theta }(S)`$ does not depend on the chosen representative $`S`$ of $`\tau `$. Thus, indeed, $`\mathrm{\Theta }`$ projects to a map from $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ to $`\text{Howe}_{}(P)`$.
To check that this map is surjective, let $`QP`$ be a holonomy-induced Howe subbundle of class $`C^0`$. Let $`\stackrel{~}{Q}Q`$ be a connected subbundle of class $`C^0`$, with structure group $`\stackrel{~}{H}`$, such that (10) holds. Due to well-known smoothing theorems \[17, Ch. I, ยง4\], $`\stackrel{~}{Q}`$ and $`Q`$ are $`C^0`$-isomorphic to $`C^{\mathrm{}}`$-subbundles of $`P`$. Hence, up to the action of $`๐ข^{C^0}`$, we may assume that $`\stackrel{~}{Q}`$ and $`Q`$ are of class $`C^{\mathrm{}}`$ themselves. Moreover, up to the action of $`G`$, $`p_0\stackrel{~}{Q}`$. Since $`M`$ is compact and $`dimM2`$, $`\stackrel{~}{Q}`$ carries a $`C^{\mathrm{}}`$-connection with holonomy group $`\stackrel{~}{H}`$ \[21, Ch. II, Thm. 8.2\]. This connection prolongs to a unique (smooth) $`A๐^k`$ obeying $`P_A=\stackrel{~}{Q}`$ and $`H_A=\stackrel{~}{H}`$. Then Lemma 3.1(a) and (10) imply
$$\mathrm{\Theta }\left(๐ข_A^{k+1}\right)=\stackrel{~}{Q}\mathrm{C}_G^2\left(\stackrel{~}{H}\right)=Q.$$
This proves surjectivity.
To show that the projected map is injective, let $`\tau ,\tau ^{}\mathrm{OT}(๐^k,๐ข^{k+1})`$. Choose representatives $`S,S^{}`$ and assume that there exist $`g๐ข^{C^0}`$, $`aG`$ such that
$$\mathrm{\Theta }\left(S^{}\right)=\vartheta _g\left(\mathrm{\Theta }\left(S\right)\right)a.$$
(11)
Consider the following lemma.
###### Lemma 3.4
Let $`A๐^k`$ and let $`QP`$ be a subbundle of class $`C^{\mathrm{}}`$. If there exists $`h๐ข^{C^0}`$ such that
$$\mathrm{\Theta }\left(๐ข_A^{k+1}\right)=\vartheta _h(Q)$$
(12)
then $`h`$ may be chosen from $`๐ข^{k+1}`$.
Before proving the lemma, let us assume that it holds and finish the arguments. Again, due to smoothing theorems, $`\mathrm{\Theta }(S)`$ is $`C^0`$-isomorphic to some $`C^{\mathrm{}}`$-subbundle $`QP`$. Then there exists $`h๐ข^{C^0}`$ such that $`\mathrm{\Theta }(S)=\vartheta _h(Q)`$. Due to Lemma 3.4, we can choose $`h๐ข^{k+1}`$. Moreover, due to (11), $`\mathrm{\Theta }(S^{})=\vartheta _{gh}\left(Qa\right)`$. By application of Lemma 3.4 again, we can achieve $`gh๐ข^{k+1}`$. This shows that we may assume, from the beginning, $`g๐ข^{k+1}`$.
Now consider (11). Since $`p_0\mathrm{\Theta }(S)`$, $`\vartheta _g(p_0)a=p_0\left(g(p_0)a\right)\mathrm{\Theta }(S^{})`$. Since also $`p_0\mathrm{\Theta }(S^{})`$, $`g(p_0)a`$ is an element of the structure group of $`\mathrm{\Theta }(S^{})`$. Then $`\mathrm{\Theta }(S^{})\left(a^1g(p_0)^1\right)=\mathrm{\Theta }(S^{})`$, so that (11) and Lemma 3.1(c) yield
$$\mathrm{\Theta }\left(S^{}\right)=\vartheta _g\left(\mathrm{\Theta }(S)\right)g(p_0)^1=\mathrm{\Theta }\left(gSg^1\right).$$
Due to Lemma 3.1(b), then $`S^{}=gSg^1`$. This proves injectivity.
Proof of Lemma 3.4: Let $`A`$ and $`Q`$ be given. Under the assumption that (12) holds, $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ and $`Q`$ have the same structure group $`H`$. There exist an open covering $`\{U_i\}`$ and local trivializations
$$\xi _i:U_i\times H\mathrm{\Theta }\left(๐ข_A^{k+1}\right)|_{U_i},\eta _i:U_i\times HQ|_{U_i}$$
of $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ and $`Q`$, respectively. These define local trivializations
$$\stackrel{~}{\xi }_i,\stackrel{~}{\eta }_i:U_i\times GP|_{U_i}$$
of $`P`$ over $`\{U_i\}`$. Here $`\eta _i`$, $`\stackrel{~}{\eta }_i`$ can be chosen from the class $`C^{\mathrm{}}`$. As for $`\xi _i`$ and $`\stackrel{~}{\xi }_i`$, we note that $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ contains the holonomy bundle $`P_A`$. A standard way of constructing local cross sections in $`P_A`$ goes as follows (cf. the proof of Lemma 1 in \[21, Ch. II,ยง7.1\]): Choose a local chart about some $`xM`$ and a point $`p`$ in the fibre of $`P_A`$ over $`x`$. Take the pre-images, under the local chart map, of straight lines running through $`x`$ and lift them to $`A`$-horizontal paths running through $`p`$. Since $`A`$ is of class $`W^k`$, the lifts are of class $`W^{k+1}`$. Hence, so is the local cross section and, therefore, the local trivialization defined by that cross section. This shows that $`\xi _i`$ and $`\stackrel{~}{\xi }_i`$ may be chosen from the class $`W^{k+1}`$.
Due to (12), the family $`\{\vartheta _h\eta _i\}`$ defines a local trivialization of class $`C^0`$ of $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ over $`\{U_i\}`$. Hence, there exists a vertical automorphism $`\vartheta ^{}`$ of class $`C^0`$ of $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ such that $`\xi _i=\vartheta ^{}\vartheta _h\eta _i`$ $`i`$. By equivariant prolongation, $`\vartheta ^{}`$ defines a unique element $`h^{}๐ข^{C^0}`$. Since $`\vartheta _h^{}`$ leaves $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)`$ invariant, $`\mathrm{\Theta }\left(๐ข_A^{k+1}\right)=\vartheta _{h^{}h}(Q)`$. Thus, by possibly redefining $`h`$ we may assume that $`h^{}=\mathrm{๐}`$, i.e., that $`\vartheta ^{}`$ is trivial. Then
$$\stackrel{~}{\xi }_i=\vartheta _h\stackrel{~}{\eta }_ii.$$
(13)
As we shall argue now, (13) implies $`h๐ข^{k+1}`$. To see this, note that $`h๐ข^{k+1}`$ iff the local representatives $`h_i=h\stackrel{~}{\eta }_i\iota `$ are of class $`W^{k+1}`$, see . Here $`\iota `$ denotes the embedding $`U_iU_i\times G`$, $`x(x,\mathrm{๐})`$. Using $`\stackrel{~}{\eta }_i(x,h_i(x))=\vartheta _h\stackrel{~}{\eta }_i(x,\mathrm{๐})`$, $`xU_i`$, we find
$$h_i=\mathrm{pr}_2\stackrel{~}{\eta }_i^1\vartheta _h\stackrel{~}{\eta }_i\iota ,$$
(14)
where $`\mathrm{pr}_2`$ is the canonical projection $`U_i\times GG`$. Inserting (13) into (14) yields
$$h_i=\mathrm{pr}_2\stackrel{~}{\eta }_i^1\stackrel{~}{\xi }_i\iota i.$$
Here $`\stackrel{~}{\xi }_i`$ is of class $`W^{k+1}`$ and the other maps are of class $`C^{\mathrm{}}`$. Thus, according to the composition rules of Sobolev mappings, $`h_i`$ is of class $`W^{k+1}`$. It follows $`h๐ข^{k+1}`$. This proves Lemma 3.4 and, therefore, Theorem 3.3.
Remarks:
1. As one important consequence of Theorem 3.3, $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ does not depend on $`k`$.
2. General arguments show that $`\text{Howe}_{}(P)`$ is countable, see \[23, ยง4.2\]. Hence, so is $`\mathrm{OT}(๐^k,๐ข^{k+1})`$. Countability of $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ is a necessary condition for this set to define a stratification in the sense of . It was first stated in Theorem 4.2.1 in . In fact, the proof of this theorem already contains most of the arguments needed to prove Theorem 3.3. Unfortunately, although in the proof of Theorem 4.2.1 the authors used that isomorphy of evolution subbundles implies conjugacy under the action of $`๐ข^{k+1}`$, they did not give an argument for that. Such an argument is provided by our Lemma 3.4.
3. Theorem 3.3 also shows that the notion of holonomy-induced Howe subbundle may be viewed as an abstract version of the notion of evolution subbundle generated by a connection.
4. The geometric ideas behind the proof of Theorem 3.3 are also contained in \[15, ยง2\]. However, a rigorous proof was not given there.
In view of Theorem 3.3, we are left with the problem of determining the set $`\text{Howe}_{}(P)`$ together with its partial ordering. This leads us to the following
Programme
Step 1 Determination of the Howe subgroups of $`G`$. Since $`G`$-action on subbundles conjugates the structure group, classification up to conjugacy is sufficient.
Step 2 Determination of Howe subbundles of $`P`$. Since subbundles are conjugate by $`๐ข^{C^0}`$ iff they are isomorphic, classification up to isomorphy is sufficient.
Step 3 Specification of the Howe subbundles which are holonomy-induced.
Step 4 Factorization by $`G`$-action
Step 5 Determination of the natural partial ordering.
In the present paper, we perform steps 1โ4 for the group $`G=\mathrm{SU}n`$. As already noted, the determination of the natural partial ordering, which includes the study of the natural partial ordering of Howe subgroups, will be published in a subsequent paper.
## 4 The Howe Subgroups of $`\mathrm{SU}n`$
Let $`\text{Howe}(\mathrm{SU}n)`$ denote the set of conjugacy classes of Howe subgroups of $`\mathrm{SU}n`$. In order to derive $`\text{Howe}(\mathrm{SU}n)`$, we consider $`\mathrm{SU}n`$ as a subset of the general linear algebra $`\mathrm{gl}(n,)`$, viewed as an associative algebra.
In the literature it is customary to consider, instead of Howe subgroups, Howe dual pairs. A Howe dual pair is an ordered pair of subgroups $`(H_1,H_2)`$ of $`G`$ such that
$$H_1=\mathrm{C}_G(H_2),H_2=\mathrm{C}_G(H_1).$$
The assignment $`H(H,\mathrm{C}_G(H))`$ defines a $`1:1`$-relation between Howe subgroups and Howe dual pairs. One also defines Howe subalgebras and Howe dual pairs in an associative algebra. We remark that, usually, one restricts attention to reductive Howe dual pairs. This means, one requires that any finite dimensional representation of the members be completely reducible. In our case this condition is automatically satisfied, because $`\mathrm{SU}n`$ is compact and Howe subgroups are always closed. Reductive Howe dual pairs play an important role in the representation theory of Lie groups. This was first observed by R. Howe . Although for $`\mathrm{SU}n`$ it is not necessary to go into the details of the classification theory of reductive Howe dual pairs, we note that there exist, essentially, two methods. One applies to the isometry groups of Hermitian spaces and uses the theory of Hermitian forms . The other method applies to complex semisimple Lie algebras and uses root space techniques .
Let $`\mathrm{K}(n)`$ denote the collection of pairs of sequences (of equal length) of positive integers
$$J=(๐ค,๐ฆ)=((k_1,\mathrm{},k_r),(m_1,\mathrm{},m_r)),r=1,2,3,\mathrm{},n,$$
which obey
$$๐ค๐ฆ=\underset{i=1}{\overset{r}{}}k_im_i=n.$$
(15)
For a given element $`J=(๐ค,๐ฆ)`$ of $`\mathrm{K}(n)`$, let $`g`$ denote the greatest common divisor of the integers $`m_1,\mathrm{},m_r`$. Define a sequence $`\stackrel{~}{๐ฆ}=(\stackrel{~}{m}_1,\mathrm{},\stackrel{~}{m}_r)`$ by $`g\stackrel{~}{m}_i=m_i`$ $`i`$. Moreover, for any permutation $`\sigma `$ of $`r`$ elements, define $`\sigma J=(\sigma ๐ค,\sigma ๐ฆ)`$.
Any $`J\mathrm{K}(n)`$ generates a canonical decomposition
$$^n=\left(^{k_1}^{m_1}\right)\mathrm{}\left(^{k_r}^{m_r}\right).$$
(16)
This decomposition, in turn, induces an injective homomorphism
$`\mathrm{gl}(k_1,)\times \mathrm{}\times \mathrm{gl}(k_r,)`$ $``$ $`\mathrm{gl}(n,)`$
$`(D_1,\mathrm{},D_r)`$ $``$ $`\left(D_1\mathrm{๐}_{m_1}\right)\mathrm{}\left(D_r\mathrm{๐}_{m_r}\right).`$ (17)
Written as a matrix w.r.t. the canonical basis, the elements of the image look like
$$\left(\begin{array}{ccc}\left(\begin{array}{cccc}D_1& 0& \mathrm{}& 0\\ 0& D_1& \mathrm{}& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& \mathrm{}& D_1\end{array}\right)& \mathrm{}& 0\\ \mathrm{}& \mathrm{}& \mathrm{}\\ 0& \mathrm{}& \left(\begin{array}{cccc}D_r& 0& \mathrm{}& 0\\ 0& D_r& \mathrm{}& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& \mathrm{}& D_r\end{array}\right)\end{array}\right)$$
($`m_1`$ times $`D_1`$, $`m_2`$ times $`D_2`$, etc.). We denote the image of this homomorphism by $`\mathrm{gl}(J,)`$, its intersection with $`\mathrm{U}n`$ by $`\mathrm{U}J`$ and its intersection with $`\mathrm{SU}n`$ by $`\mathrm{SU}J`$. Note that $`\mathrm{U}J`$ is the image of the restriction of (17) to $`\mathrm{U}k_1\times \mathrm{}\times \mathrm{U}k_r`$.
###### Lemma 4.1
(a) A $``$-subalgebra of $`\mathrm{gl}(n,)`$ is Howe if and only if it is conjugate, under the action of $`\mathrm{SU}n`$ by inner automorphisms, to $`\mathrm{gl}(J,)`$ for some $`J\mathrm{K}(n)`$.
(b) A subgroup of $`\mathrm{U}n`$ (resp. $`\mathrm{SU}n`$) is Howe if and only if it is conjugate, under the action of $`\mathrm{SU}n`$ by inner automorphisms, to $`\mathrm{U}J`$ (resp. $`\mathrm{SU}J`$) for some $`J\mathrm{K}(n)`$.
Remark: Assertion (a) is a version of the structure theorem for finite-dimensional von Neumann algebras.
Proof: (a) For any $`J\mathrm{K}(n)`$, $`\mathrm{gl}(J,)`$ is a $``$-subalgebra of $`\mathrm{gl}(n,)`$ containing $`\mathrm{๐}_n`$. Thus, $`\mathrm{gl}(J,)`$ is a von Neumann algebra and the Double Commutant Theorem says that $`\mathrm{gl}(J,)`$ is Howe. Conversely, let $`L`$ be a Howe $``$-subalgebra of $`\mathrm{gl}(n,)`$. Let $`L^{}`$ denote the centralizer of $`L`$ in $`\mathrm{gl}(n,)`$ and let $`\stackrel{~}{L}`$ denote the subalgebra generated by $`L`$ and $`L^{}`$. Decompose
$$^n=V_1\mathrm{}V_r$$
(18)
into mutually orthogonal, $`\stackrel{~}{L}`$-irreducible subspaces. Due to Schurโs Lemma, for each $`i`$,
$$V_i=W_i^{m_i},$$
(19)
where $`W_i`$ is $`L`$-irreducible and $`m_i`$ is a positive integer. Moreover,
$`L^{}|_{V_i}`$ $``$ $`\left\{\mathrm{id}_{W_i}D_i^{}\right|D_i^{}\mathrm{gl}(m_i,)\},`$ (20)
$`L|_{V_i}`$ $``$ $`\left\{D_i\mathrm{๐}_{m_i}\right|D_i\text{End}(W_i)\}.`$ (21)
Since $`L^{}`$ is the full centralizer of $`L`$, one has equality in (20). As a consequence, since $`L`$ is the full centralizer of $`L^{}`$, there holds equality in (21). Now let $`k_i=dimW_i`$ and $`J=(๐ค,๐ฆ)`$. By identifying $`W_i^{k_i}`$, we arrive at a decomposition of $`^n`$ of the form (16). There exists $`T\mathrm{SU}n`$ transforming this decomposition into the canonical one induced by $`J`$. Then the inner automorphism of $`\mathrm{gl}(n,)`$ defined by $`T`$ transforms $`L`$ into $`\mathrm{gl}(J,)`$.
(b) We give only the proof for $`\mathrm{SU}n`$. Let $`H`$ be a Howe subgroup of $`\mathrm{SU}n`$. Then
$$H=\mathrm{C}_{\mathrm{SU}n}(K)=\mathrm{C}_{\mathrm{gl}(n,)}(K)\mathrm{SU}n$$
for some subset $`K\mathrm{SU}n`$. By construction, $`\mathrm{C}_{\mathrm{gl}(n,)}(K)`$ is a Howe $``$-subalgebra of $`\mathrm{gl}(n,)`$. Hence, by virtue of (a), it is conjugate, under $`\mathrm{SU}n`$-action, to $`\mathrm{gl}(J,)`$ for some $`J`$. Then $`H`$ is conjugate, in $`\mathrm{SU}n`$, to $`\mathrm{gl}(J,)\mathrm{SU}n=\mathrm{SU}J`$. Conversely, let $`J\mathrm{K}(n)`$. Due to (a), there exists $`J^{}\mathrm{K}(n)`$ and $`T\mathrm{SU}n`$ such that $`\mathrm{gl}(J,)=\mathrm{C}_{\mathrm{gl}(n,)}\left(T\mathrm{gl}(J^{},)T^1\right)`$. Taking into account that $`\left(T\mathrm{gl}(J^{},)T^1\right)\mathrm{SU}n`$ spans $`\left(T\mathrm{gl}(J^{},)T^1\right)`$, we obtain
$$\begin{array}{ccc}\hfill \mathrm{SU}J& =& \mathrm{gl}(J,)\mathrm{SU}n\hfill \\ & =& \mathrm{C}_{\mathrm{gl}(n,)}\left(T\mathrm{gl}(J^{},)T^1\mathrm{SU}n\right)\mathrm{SU}n\hfill \\ & =& \mathrm{C}_{\mathrm{SU}n}\left(T\mathrm{gl}(J^{},)T^1\mathrm{SU}n\right).\hfill \end{array}$$
This shows that $`\mathrm{SU}J`$ is Howe.
###### Lemma 4.2
Let $`J,J^{}\mathrm{K}(n)`$. Then $`\mathrm{gl}(J,)`$ and $`\mathrm{gl}(J^{},)`$ are conjugate under the action of $`\mathrm{SU}n`$ by inner automorphisms if and only if there exists a permutation $`\sigma `$ of $`1,\mathrm{},r`$ such that $`J^{}=\sigma J`$.
Proof: Let $`T\mathrm{SU}n`$ such that $`\mathrm{gl}(J^{},)=T^1\mathrm{gl}(J,)T`$. Then $`\mathrm{gl}(J,)`$ and $`\mathrm{gl}(J^{},)`$ are isomorphic. Hence, there exists a permutation $`\sigma `$ of $`1,\mathrm{},r`$ such that $`๐ค^{}=\sigma ๐ค`$. Moreover, $`T`$ is an isomorphism of the representations
$$\begin{array}{c}\mathrm{gl}(k_1,)\times \mathrm{}\times \mathrm{gl}(k_r,)\stackrel{J}{}\mathrm{gl}(n,),\hfill \\ \mathrm{gl}(k_1,)\times \mathrm{}\times \mathrm{gl}(k_r,)\stackrel{\sigma }{}\mathrm{gl}(k_{\sigma (1)},)\times \mathrm{}\times \mathrm{gl}(k_{\sigma (r)},)\stackrel{J^{}}{}\mathrm{gl}(n,),\hfill \end{array}$$
where $`J`$, $`J^{}`$ indicate the embeddings (17) defined by $`J`$ and $`J^{}`$, respectively. In particular, it does not change the multiplicities of the irreducible factors. This implies $`๐ฆ^{}=\sigma ๐ฆ`$. It follows $`J^{}=\sigma J`$.
Conversely, assume that $`J^{}=\sigma J`$ for some permutation $`\sigma `$ of $`1,\mathrm{},r`$. Consider the canonical decompositions (16) defined by $`J`$ and $`J^{}`$, respectively. There exists $`T\mathrm{SU}n`$ which maps the factors $`^{k_i^{}}^{l_i^{}}`$ of the second decomposition identically onto the factors $`^{k_{\sigma (i)}}^{m_{\sigma (i)}}`$ of the first one, for $`i=1,\mathrm{},r`$. It is not difficult to see that $`\mathrm{gl}(J^{},)=T^1\mathrm{gl}(J,)T`$.
As a consequence of Lemma 4.2, we introduce an equivalence relation on the set $`\mathrm{K}(n)`$ by writing $`JJ^{}`$ iff there exists a permutation of $`1,\mathrm{},r`$ such that $`J^{}=\sigma J`$. Let $`\widehat{\mathrm{K}}(n)`$ denote the set of equivalence classes.
###### Theorem 4.3
The assignment $`J\mathrm{SU}J`$ induces a bijection from $`\widehat{\mathrm{K}}(n)`$ onto $`\text{Howe}(\mathrm{SU}n)`$.
Proof: According to Lemma 4.1, the assignment $`J\mathrm{SU}J`$ induces a surjective map $`\mathrm{K}(n)\text{Howe}(\mathrm{SU}n)`$. Due to Lemma 4.2, this map projects to $`\widehat{\mathrm{K}}(n)`$ and the projected map is injective.
This concludes the classification of Howe subgroups of $`\mathrm{SU}n`$, i.e., Step $`1`$ of our programme.
## 5 The Howe Subbundles of $`\mathrm{SU}n`$-Bundles
In this chapter, let $`P`$ be a principal $`\mathrm{SU}n`$-bundle over $`M`$, $`dimM4`$. We are going to derive the Howe subbundles of $`P`$ up to isomorphy. Since the action of $`\mathrm{SU}n`$ on subbundles conjugates the structure group, it suffices to consider reductions of $`P`$ to the subgroups $`\mathrm{SU}J`$, $`J\mathrm{K}(n)`$. Thus, let $`J\mathrm{K}(n)`$ be fixed. Let $`\text{Bun}(M,\mathrm{SU}J)`$ denote the set of isomorphism classes of principal $`\mathrm{SU}J`$-bundles over $`M`$ (where principal bundle isomorphisms are assumed to commute with the structure group action and to project to the identical map on the base space). Moreover, let $`\text{Red}(P,\mathrm{SU}J)`$ denote the set of isomorphism classes of reductions of $`P`$ to the subgroup $`\mathrm{SU}J\mathrm{SU}n`$.
We shall first derive a description of $`\text{Bun}(M,\mathrm{SU}J)`$ in terms of suitable characteristic classes and then give a characterization of the subset $`\text{Red}(P,\mathrm{SU}J)`$. The classification of $`\text{Bun}(M,\mathrm{SU}J)`$ will be performed by constructing the Postnikov tower of the classifying space $`\mathrm{BSU}J`$ up to level $`5`$. For the convenience of the reader, the basics of this method will be briefly explained below.
Note that in the sequel maps of topological spaces are always assumed to be continuous, without explicitly stating this.
### 5.1 Preliminaries
#### Universal Bundles and Classifying Spaces
Let $`G`$ be a Lie group. As a basic fact in bundle theory, there exists a so-called universal $`G`$-bundle
$$G\mathrm{E}G\mathrm{B}G$$
(22)
with the following property: For any CW-complex (hence, in particular, any manifold) $`X`$ the assignment
$$[X,\mathrm{B}G]\text{Bun}(X,G),ff^{}\mathrm{E}G,$$
(23)
is a bijection . Here $`[,]`$ means the set of homotopy classes of maps and $`f^{}`$ denotes the pull-back of bundles. Both $`\mathrm{E}G`$ and $`\mathrm{B}G`$ can be realized as CW-complexes. They are unique up to homotopy equivalence. $`\mathrm{B}G`$ is called the classifying space of $`G`$. The homotopy class of maps $`X\mathrm{B}G`$ associated to $`P\text{Bun}(X,G)`$ by virtue of (23) is called the classifying map of $`P`$. We denote it by $`f_P`$. Note that a principal $`G`$-bundle is universal iff its total space is contractible. In particular,
$$\pi _i(G)\pi _{i+1}(\mathrm{B}G),i=0,1,2,\mathrm{}.$$
(24)
This is an immediate consequence of the exact homotopy sequence of fibre spaces .
#### Associated Principal Bundles Defined by Homomorphisms
Let $`\phi :GG^{}`$ be a Lie group homomorphism and let $`P\text{Bun}(X,G)`$. By virtue of the action
$$G\times G^{}G^{},(a,a^{})\phi (a)a^{},$$
$`G^{}`$ becomes a left $`G`$-space and we have an associated bundle $`P^{[\phi ]}=P\times _GG^{}`$. Observe that $`P^{[\phi ]}`$ can be viewed as a principal $`G^{}`$-bundle with right $`G^{}`$-action
$$P^{[\phi ]}\times G^{}P^{[\phi ]},([(p,a^{})],b^{})[(p,a^{}b^{})].$$
One has the natural bundle morphism
$$\psi :PP^{[\phi ]},p[(p,\mathrm{๐}_G^{})].$$
(25)
It obeys $`\psi (pa)=\psi (p)\phi (a)`$ and projects to the identical map on $`X`$.
In the special case where $`\phi `$ is a Lie subgroup embedding, the natural bundle morphism (25) is an embedding of $`P`$ onto a subbundle of $`P^{[\phi ]}`$. In this case, if no confusion about $`\phi `$ can arise, we shall often write $`P^{[G^{}]}`$ instead of $`P^{[\phi ]}`$. Note that $`P^{[\phi ]}`$ is the extension of $`P`$ by $`G^{}`$ and $`P`$ is a reduction of $`P^{[\phi ]}`$ to the subgroup $`(G,\phi )`$.
#### Classifying Maps Associated to Homomorphisms
Again, let $`\phi :GG^{}`$ be a homomorphism. There exists a map $`\mathrm{B}\phi :\mathrm{B}G\mathrm{B}G^{}`$, associated to $`\phi `$, which is defined as the classifying map of the principal $`G^{}`$-bundle $`(\mathrm{E}G)^{[\phi ]}`$ associated to the universal $`G`$-bundle $`\mathrm{E}G`$. It has the following functorial property: For $`\phi :GG^{}`$ and $`\phi ^{}:G^{}G^{\prime \prime }`$ there holds
$$\mathrm{B}(\psi \phi )=\mathrm{B}\psi \mathrm{B}\phi .$$
(26)
Using $`\mathrm{B}\phi `$, the classifying map of $`P^{[\phi ]}`$ can be expressed through that of $`P`$:
$$f_{P^{[\phi ]}}=\mathrm{B}\phi f_P.$$
(27)
In the special case where $`\phi `$ is a normal Lie subgroup embedding, the short exact sequence of Lie group homomorphisms
$$\mathrm{๐}G\stackrel{\phi }{}G^{}\stackrel{p}{}G/G^{}\mathrm{๐}$$
(28)
induces a principal bundle
$$G^{}/G\mathrm{B}G\stackrel{\mathrm{B}\phi }{}\mathrm{B}G^{}.$$
(29)
The classifying map of this bundle is $`\mathrm{B}p`$ , where $`p`$ denotes the natural projection.
#### Characteristic Classes
Let $`G`$ be a Lie group. Consider the cohomology ring $`H^{}(\mathrm{B}G,\pi )`$ of the classifying space with values in some Abelian group $`\pi `$. For any $`P\text{Bun}(X,G)`$, the homomorphism $`\left(f_P\right)^{}`$, induced on cohomology, maps $`H^{}(\mathrm{B}G,\pi )`$ to $`H^{}(X,\pi )`$. Therefore, given $`\gamma H^{}(\mathrm{B}G,\pi )`$, one can define a map
$$\chi _\gamma :\text{Bun}(X,G)H^{}(X,\pi ),P\left(f_P\right)^{}\gamma .$$
(30)
This is called the characteristic class for $`G`$-bundles over $`X`$ defined by $`\gamma `$. By construction, one has the following universal property of characteristic classes: Let $`f:XX^{}`$ be a map and let $`P^{}\text{Bun}(X^{},G)`$. Then
$$\chi _\gamma \left(f^{}P^{}\right)=f^{}\chi _\gamma \left(P^{}\right).$$
(31)
Observe that if two bundles are isomorphic then their images under arbitrary characteristic classes coincide, whereas the converse, in general, does not hold. This is due to the fact that characteristic classes can control maps $`X\mathrm{B}G`$ only on the level of the homomorphisms induced on cohomology. In general, the latter do not give sufficient information on the homotopy properties of the maps. In certain cases, however, they do. For example, such cases are obtained by specifying $`G`$ to be $`\mathrm{U1}`$ or discrete, or by restricting $`X`$ in dimension. In these cases there exist sets of characteristic classes which classify $`\text{Bun}(X,G)`$. In the sequel we will utilize this to determine $`\text{Bun}(M,\mathrm{SU}J)`$.
#### Eilenberg-MacLane Spaces
Let $`\pi `$ be a group and $`n`$ a positive integer. An arcwise connected CW-complex $`X`$ is called an Eilenberg-MacLane space of type $`K(\pi ,n)`$ iff $`\pi _n(X)=\pi `$ and $`\pi _i(X)=0`$ for $`in`$. Eilenberg-MacLane spaces exist for any choice of $`\pi `$ and $`n`$, provided $`\pi `$ is commutative for $`n2`$. They are unique up to homotopy equivalence.
The simplest example of an Eilenberg-MacLane space is the $`1`$-sphere $`\mathrm{S}^1`$, which is of type $`K(,1)`$. Two further examples, $`K(,2)`$ and $`K(_g,1)`$, are briefly discussed in Appendix A. Note that Eilenberg-MacLane spaces are, apart from very special examples, infinite dimensional. Also note that, up to homotopy equivalence one has
$$K(\pi _1\times \pi _2,n)=K(\pi _1,n)\times K(\pi _2,n).$$
(32)
Now assume $`\pi `$ to be commutative also in the case $`n=1`$. Due to the Universal Coefficient Theorem, $`\text{Hom}(H_n(K(\pi ,n)),\pi )`$ is isomorphic to a subgroup of $`H^n(K(\pi ,n),\pi )`$. Due to the Hurewicz Theorem, $`H_n(K(\pi ,n))\pi _n(K(\pi ,n))=\pi `$. It follows that $`H^n(K(\pi ,n),\pi )`$ contains elements which correspond to isomorphisms $`H_n(K(\pi ,n))\pi `$. Such elements are called characteristic. If $`\gamma H^n(K(\pi ,n),\pi )`$ is characteristic then for any CW-complex $`X`$, the map
$$[X,K(\pi ,n)]H^n(X,\pi ),ff^{}\gamma ,$$
(33)
is a bijection \[8, ยงVII.12\]. In this sense, Eilenberg-MacLane spaces provide a link between homotopy properties and cohomology. We remark that the bijection (33) induces an Abelian group structure on the set $`[X,K(\pi ,n)]`$.
#### Path-Loop-Fibration
Let $`X`$ be an arcwise connected topological space. Consider the path-loop fibration over $`X`$
$$\mathrm{\Omega }(X)P(X)X,$$
(34)
where $`\mathrm{\Omega }(X)`$ and $`P(X)`$ denote the loop space and the path space of $`X`$, respectively (both based at some point $`x_0X`$). Since $`P(X)`$ is contractible, the exact homotopy sequence induced by the fibration (34) implies $`\pi _i(\mathrm{\Omega }(X))\pi _{i+1}(X)`$, $`i=0,1,2,\mathrm{}`$. Thus, for $`X=K(\pi ,n)`$,
$$\pi _i\left(\mathrm{\Omega }(K(\pi ,n+1))\right)\pi _{i+1}\left(K(\pi ,n+1)\right)=\{\begin{array}{ccc}\pi & |\hfill & i=n\hfill \\ 0& |\hfill & \text{otherwise}\hfill \end{array}$$
Hence, $`\mathrm{\Omega }(K(\pi ,n+1))=K(\pi ,n)`$ $`n`$ and the path-loop fibration over $`K(\pi ,n+1)`$ reads
$$K(\pi ,n)P(K(\pi ,n+1))K(\pi ,n+1).$$
(35)
#### Postnikov Tower
A map $`f:XX^{}`$ of topological spaces is called an $`n`$-equivalence iff the homomorphism induced on homotopy groups $`f_{}:\pi _i(X)\pi _i(X^{})`$ is an isomorphism for $`i<n`$ and surjective for $`i=n`$. One also defines the notion of an $`\mathrm{}`$-equivalence, which is often called weak homotopy equivalence.
Let $`f:XX^{}`$ be an $`n`$-equivalence and let $`Y`$ be a CW-complex. Then the map $`[Y,X][Y,X^{}]`$, $`gfg`$, is bijective for $`dimY<n`$ and surjective for $`dimY=n`$ \[8, Ch. VII, Cor. 11.13\].
A CW-complex $`Y`$ is called $`n`$-simple iff it is arcwise connected and the action of $`\pi _1(Y)`$ on $`\pi _i(Y)`$ is trivial for $`1in`$. It is called simple iff it is $`n`$-simple for all $`n`$.
The following theorem describes how a simple CW-complex can be approximated by $`n`$-equivalent spaces constructed from Eilenberg-MacLane spaces.
###### Theorem 5.1
Let $`Y`$ be a simple CW-complex. There exist
(a) a sequence of CW-complexes $`Y_n`$ and principal fibrations
$$K(\pi _n(Y),n)Y_{n+1}\stackrel{q_n}{}Y_n,n=1,2,3,\mathrm{},$$
(36)
induced by maps $`\theta _n:Y_nK(\pi _n(Y),n+1)`$,
(b) a sequence of $`n`$-equivalences $`y_n:YY_n`$, $`n=1,2,3,\mathrm{}`$,
such that $`Y_1=`$ (one point space) and $`q_ny_{n+1}=y_n`$ for all $`n`$.
Proof: The assumption that $`Y`$ be a simple CW-complex implies that the constant map $`Y`$ is a simple map (see \[8, Ch. VII, Def. 13.4\] for a definition of the latter). Thus, the assertion is a consequence of a more general theorem about simple maps given in \[8, Ch. VII, Thm. 13.7\].
Remarks:
1. The sequence of spaces and maps $`(Y_n,y_n,q_n)`$, $`n=1,2,3,\mathrm{}`$, is called a Postnikov tower, or Postnikov system, or Postnikov decomposition of $`Y`$.
2. For the principal fibrations (36) to be induced by a map $`\theta _n:Y_nK(\pi _n(Y),n+1)`$ means that they are given as pull-back of the path-loop fibration (35) over $`K(\pi _n(Y),n+1)`$.
The theorem allows one to successively construct $`Y_n`$ for given $`n`$, starting from $`Y_1=`$. For example, for $`n=5`$ such constructions have been carried out for $`Y=\mathrm{BU}k`$ in \[5, ยง4.2\], or for $`Y=\mathrm{BPU}k`$, where $`\mathrm{PU}k`$ denotes the projective unitary group, in . In the sequel, we shall construct $`\left(\mathrm{BSU}J\right)_5`$. For this purpose, we need information about the low-dimensional homotopy groups of $`\mathrm{SU}J`$.
### 5.2 The Homotopy Groups of $`\mathrm{SU}J`$
For $`a`$ being a positive integer, denote the canonical embedding $`_a\mathrm{U1}`$ by $`j_a`$ and the endomorphism of $`\mathrm{U1}`$ mapping $`zz^a`$ by $`p_a`$. Moreover, let $`j_J`$ and $`i_J`$ denote the natural embeddings $`\mathrm{SU}J\mathrm{U}J`$ and $`\mathrm{U}J\mathrm{U}n`$, respectively.
Using the natural projections $`\mathrm{pr}_i^{\mathrm{U}J}:\mathrm{U}J\mathrm{U}k_i`$, we define a homomorphism
$$\lambda _J:\mathrm{U}J\mathrm{U1},D_{i=1}^rp_{\stackrel{~}{m}_i}det_{\mathrm{U}k_i}\mathrm{pr}_i^{\mathrm{U}J}(D).$$
(37)
The following diagram commutes:
(38)
Accordingly, the restriction of $`\lambda _J`$ to the subgroup $`\mathrm{SU}J`$ takes values in $`j_g(_g)\mathrm{U1}`$. Thus, we can define a homomorphism $`\lambda _J^\mathrm{S}:\mathrm{SU}J_g`$ by the following commutative diagram:
(39)
Let $`(\mathrm{SU}J)_0`$ denote the arcwise connected component of the identity. Note that it is also a connected component, because $`\mathrm{SU}J`$ is a closed subgroup of $`\mathrm{GL}(n,)`$.
###### Lemma 5.2
The homomorphism $`\lambda _J^\mathrm{S}`$ projects to an isomorphism $`\mathrm{SU}J/(\mathrm{SU}J)_0_g`$.
Proof: Consider the homomorphism $`\lambda _J^\mathrm{S}:\mathrm{SU}J_g`$. The target space being discrete, $`\lambda _J^\mathrm{S}`$ must be constant on connected components. Hence $`(\mathrm{SU}J)_0\mathrm{ker}\lambda _J^\mathrm{S}`$, so that $`\lambda _J^\mathrm{S}`$ projects to a homomorphism $`\mathrm{SU}J/(\mathrm{SU}J)_0_g`$. The latter is surjective, because $`\lambda _J^\mathrm{S}`$ is surjective. To prove injectivity, we show $`\mathrm{ker}\lambda _J^\mathrm{S}(\mathrm{SU}J)_0`$. Let $`D\mathrm{ker}\lambda _J^\mathrm{S}`$ and denote $`D_i=\mathrm{pr}_i^{\mathrm{U}J}j_J(D)`$. Define a homomorphism
$$\phi :\mathrm{U1}^r\mathrm{U1},(z_1,\mathrm{},z_r)z_1^{\stackrel{~}{m}_1}\mathrm{}z_r^{\stackrel{~}{m}_r}.$$
Then
$$\lambda _J^\mathrm{S}(D)=\phi (det_{\mathrm{U}k_1}D_1,\mathrm{},det_{\mathrm{U}k_r}D_r).$$
By assumption, $`(det_{\mathrm{U}k_1}D_1,\mathrm{},det_{\mathrm{U}k_r}D_r)\mathrm{ker}\phi `$. Since the exponents defining $`\phi `$ have greatest common divisor 1, $`\mathrm{ker}\phi `$ is connected. Thus, there exists a path $`(\gamma _1(t),\mathrm{},\gamma _r(t))`$ in $`\mathrm{ker}\phi `$ running from $`(det_{\mathrm{U}k_1}D_1,\mathrm{},det_{\mathrm{U}k_r}D_r)`$ to $`(1,\mathrm{},1)`$. For each $`i=1,\mathrm{},r`$, define a path $`G_i(t)`$ in $`\mathrm{U}k_i`$ as follows: First, go from $`D_i`$ to $`(det_{\mathrm{U}k_i}D_i)\mathrm{๐}_{k_i1}`$, keeping the determinant constant, thus using connectedness of $`\mathrm{SU}k_i`$. Next, use the path $`\gamma _i(t)\mathrm{๐}_{k_i1}`$ to get to $`\mathrm{๐}_{k_i}`$. By construction, the image of $`(G_1(t),\mathrm{},G_r(t))`$ under the embedding (17) is a path in $`\mathrm{SU}J`$ connecting $`D`$ with $`\mathrm{๐}_n`$. This proves $`\mathrm{ker}\lambda _J^\mathrm{S}(\mathrm{SU}J)_0`$.
###### Theorem 5.3
The homotopy groups of $`\mathrm{SU}J`$ are
$$\pi _i(\mathrm{SU}J)\{\begin{array}{ccc}_g& |& i=0\hfill \\ ^{(r1)}& |& i=1\hfill \\ \pi _i(\mathrm{U}k_1)\mathrm{}\pi _i(\mathrm{U}k_r)& |& i>1.\hfill \end{array}$$
In particular, $`\pi _1(\mathrm{SU}J)`$ and $`\pi _3(\mathrm{SU}J)`$ are torsion-free.
Proof: For $`i=0`$, $`\pi _0(\mathrm{SU}J)=\mathrm{SU}J/(\mathrm{SU}J)_0`$. This group is given by Lemma 5.2. For $`i>1`$, the assertion follows immediately from the exact homotopy sequence induced by the bundle $`\mathrm{SU}J\mathrm{U}J\stackrel{det_{\mathrm{U}n}}{}\mathrm{U1}`$. For $`i=1`$, consider the following portion of this sequence:
$$\begin{array}{ccccccccccc}\pi _2(\mathrm{U1})& & \pi _1(\mathrm{SU}J)& & \pi _1(\mathrm{U}J)& \stackrel{(det_{\mathrm{U}n})_{}}{}& \pi _1(\mathrm{U1})& & \pi _0(\mathrm{SU}J)& & \pi _0(\mathrm{U}J)\\ 0& & \pi _1(\mathrm{SU}J)& & ^r& & & & _g& & 0.\end{array}$$
One has $`^r/\mathrm{ker}((det_{\mathrm{U}n})_{})\mathrm{im}((det_{\mathrm{U}n})_{})`$. Exactness implies $`\mathrm{ker}((det_{\mathrm{U}n})_{})\pi _1(\mathrm{SU}J)`$ and $`\mathrm{im}((det_{\mathrm{U}n})_{})=g`$. It follows $`\pi _1(\mathrm{SU}J)^{(r1)}`$, as asserted.
### 5.3 The Postnikov Tower of $`\mathrm{BSU}J`$ up to Level 5
Let $`r^{}`$ denote the number of indices $`i`$ for which $`k_i>1`$.
###### Theorem 5.4
The $`5`$th level of the Postnikov tower of $`\mathrm{BSU}J`$ is given by
$$(\mathrm{BSU}J)_5=K(_g,1)\times \underset{j=1}{\overset{r1}{}}K(,2)\times \underset{j=1}{\overset{r^{}}{}}K(,4).$$
(40)
Proof: First, we check that $`\mathrm{BSU}J`$ is a simple space. To see this, note that any inner automorphism of $`\mathrm{SU}J`$ is generated by an element of $`(\mathrm{SU}J)_0`$, hence is homotopic to the identity automorphism. Consequently, the natural action of $`\pi _0(\mathrm{SU}J)`$ on $`\pi _{i1}(\mathrm{SU}J)`$, $`i=1,2,3,\mathrm{}`$ , induced by inner automorphisms, is trivial. Since the natural isomorphisms $`\pi _{i1}(\mathrm{SU}J)\pi _i(\mathrm{BSU}J)`$ transform this action into that of $`\pi _1(\mathrm{BSU}J)`$ on $`\pi _i(\mathrm{BSU}J)`$, the latter is trivial, too. Thus, we can apply Theorem 5.1 to construct the Postnikov tower of $`\mathrm{BSU}J`$ up to level $`5`$. According to Theorem 5.3, the relevant homotopy groups are
$$\pi _1(\mathrm{BSU}J)=_g,\pi _2(\mathrm{BSU}J)=^{(r1)},\pi _3(\mathrm{BSU}J)=0,\pi _4(\mathrm{BSU}J)=^r^{}.$$
(41)
Moreover, we note that $`H^{}(K(,2),)`$ is torsion-free, and that
$$H^{2i+1}(K(,2),)=0,H^{2i+1}(K(_g,1),)=0,i=0,1,2,\mathrm{},$$
(42)
see Appendix A. We start with $`(\mathrm{BSU}J)_1=`$.
$`(\mathrm{BSU}J)_2:`$ Being a fibration over $`(\mathrm{BSU}J)_1`$, $`(\mathrm{BSU}J)_2`$ must coincide with the fibre:
$$(\mathrm{BSU}J)_2=K(_g,1).$$
(43)
$`(\mathrm{BSU}J)_3:`$ In view of (43) and (41), $`(\mathrm{BSU}J)_3`$ is the total space of a fibration
$$K(^{(r1)},2)(\mathrm{BSU}J)_3\stackrel{q_2}{}K(_g,1)$$
(44)
induced from the path-loop fibration over $`K(^{(r1)},3)`$ by some map $`\theta _2:K(_g,1)K(^{(r1)},3)`$. Note that $`K(^{(r1)},n)=_{j=1}^{r1}K(,n)`$ $`n`$. Then, due to (33),
$$[K(_g,1),K(^{(r1)},3)]=\underset{i=1}{\overset{r1}{}}H^3(K(_g,1),).$$
Here the rhs. is trivial by (42). Hence, $`\theta _2`$ is homotopic to a constant map, so that the fibration (44) is trivial. It follows
$$(\mathrm{BSU}J)_3=K(_g,1)\times \underset{j=1}{\overset{r1}{}}K(,2).$$
(45)
$`(\mathrm{BSU}J)_4:`$ In view of (41), $`(\mathrm{BSU}J)_4`$ is given by a fibration over $`(\mathrm{BSU}J)_3`$ with fibre $`K(0,3)=`$. Hence, it just coincides with the base space.
$`(\mathrm{BSU}J)_5:`$ According to (41) and (45), $`(\mathrm{BSU}J)_5`$ is the total space of a fibration
$$K(^r^{},4)(\mathrm{BSU}J)_5\stackrel{q_4}{}K(_g,1)\times \underset{j=1}{\overset{r1}{}}K(,2),$$
(46)
which is induced by a map $`\theta _4`$ from the base to $`K(^r^{},5)`$. Similarly to the case of $`\theta _2`$,
$$[K(_g,1)\times \underset{j=1}{\overset{r1}{}}K(,2),K(^r^{},5)]=\underset{i=1}{\overset{r^{}}{}}H^5(K(_g,1)\times \underset{j=1}{\overset{r1}{}}K(,2),).$$
(47)
Since $`H^{}(K(,2),)`$ is torsion-free, we can apply the Kรผnneth Theorem for cohomology \[25, Ch. XIII, Cor. 11.3\] to obtain
$`H^5(K(_g,1)\times {\displaystyle \underset{j=1}{\overset{r1}{}}}K(,2),)`$
$`{\displaystyle \underset{\genfrac{}{}{0pt}{}{j,j_1,\mathrm{},j_{r\text{-}1}}{j+j_1+\mathrm{}+j_{r1}=5}}{}}H^j(K(_g,1),)H^{j_1}(K(,2),)\mathrm{}H^{j_{r1}}(K(,2),).`$
By (42), each summand of the rhs. is trivial, because it contains a tensor factor of odd degree. Hence, (47) is trivial, and so is the fibration (46). This proves the assertion.
The fact that $`(\mathrm{BSU}J)_5`$ is a direct product of Eilenberg-MacLane spaces immediately yields the following corollary.
###### Corollary 5.5
Let $`J\mathrm{K}(n)`$ and $`dimM4`$. Let $`P,P^{}\text{Bun}(M,\mathrm{SU}J)`$. Assume that for any characteristic class $`\alpha `$ defined by an element of $`H^1(\mathrm{BSU}J,_g)`$, $`H^2(\mathrm{BSU}J,)`$, or $`H^4(\mathrm{BSU}J,)`$ there holds $`\alpha (P)=\alpha (P^{})`$. Then $`P`$ and $`P^{}`$ are isomorphic.
Proof: Let $`\mathrm{pr}_1`$, $`\mathrm{pr}_{21},\mathrm{},\mathrm{pr}_{2r\text{-}1}`$, and $`\mathrm{pr}_{41},\mathrm{},\mathrm{pr}_{4r^{}}`$ denote the natural projections of the direct product $`K(_g,1)\times _{i=1}^{r1}K(,2)\times _{i=1}^r^{}K(,4)`$ onto its factors. Let $`\gamma _1`$, $`\gamma _2`$, and $`\gamma _4`$ be characteristic elements of $`H^1(K(_g,1),_g)`$, $`H^2(K(,2),)`$, and $`H^4(K(,4),)`$, respectively. Consider the map
$$\begin{array}{ccc}\hfill \phi :[M,\mathrm{BSU}J]& & [M,(\mathrm{BSU}J)_5]\hfill \\ & & [M,K(_g,1)]\times _{i=1}^{r1}[M,K(,2)]\times _{i=1}^r^{}[M,K(,4)]\hfill \\ & & H^1(M,_g)\times _{i=1}^{r1}H^2(M,)\times _{i=1}^r^{}H^4(M,),\hfill \\ \hfill f& & (f^{}(\mathrm{pr}_1y_5)^{}\gamma _1,\{f^{}(\mathrm{pr}_{2i}y_5)^{}\gamma _2\}_{i=1}^{r1},\{f^{}(\mathrm{pr}_{4i}y_5)^{}\gamma _4\}_{i=1}^r^{}),\hfill \end{array}$$
(48)
where $`y_5:\mathrm{BSU}J(\mathrm{BSU}J)_5`$ is the $`5`$-equivalence provided by Theorem 5.1. According to Theorem 5.4, the second step of $`\phi `$ and, therefore, the whole map, is a bijection.
Now let $`P,P^{}\text{Bun}(M,\mathrm{SU}J)`$ as proposed in the assertion. Then, by assumption, the homomorphisms $`\left(f_P\right)^{}`$ and $`\left(f_P^{}\right)^{}`$, induced on $`H^1(\mathrm{BSU}J,_g)`$, $`H^2(\mathrm{BSU}J,)`$, and $`H^4(\mathrm{BSU}J,)`$, coincide. This implies $`\phi \left(f_P\right)=\phi \left(f_P^{}\right)`$. Hence, $`f_P`$ and $`f_P^{}`$ are homotopic. This proves the corollary.
We remark that, of course, the cohomology elements $`\left(\mathrm{pr}_1y_5\right)^{}\gamma _1`$, $`\left(\mathrm{pr}_{2i}y_5\right)^{}\gamma _2`$, $`i=1,\mathrm{},r1`$, and $`\left(\mathrm{pr}_{4i}y_5\right)^{}\gamma _4`$, $`i=1,\mathrm{},r^{}`$ define a set of characteristic classes which classifies $`\text{Bun}(M,\mathrm{SU}J)`$. These classes are independent and surjective. However, they are hard to handle, because we do not know the homomorphism $`y_5^{}`$ explicitly. Therefore, we prefer to work with characteristic classes defined by some natural generators of the cohomology groups in question. The price we have to pay for this is that the classes so constructed are subject to a relation and that we have to determine their image explicitly.
### 5.4 Generators for $`H^{}(\mathrm{BSU}J,)`$
Instead of generators for the groups $`H^2(\mathrm{BU}J,)`$ and $`H^4(\mathrm{BU}J,)`$ only, we can construct generators for the whole cohomology algebra $`H^{}(\mathrm{BSU}J,)`$ without any additional effort.
Consider the homomorphisms and induced homomorphisms
$$\begin{array}{ccccc}\hfill \mathrm{SU}J& \stackrel{j_J}{}& \mathrm{U}J& \stackrel{\mathrm{pr}_i^{\mathrm{U}J}}{}& \mathrm{U}k_i,\hfill \\ \hfill H^{}(\mathrm{BSU}J,)& \stackrel{\left(\mathrm{B}j_J\right)^{}}{}& H^{}(\mathrm{BU}J,)& \stackrel{\left(\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}}{}& H^{}(\mathrm{BU}k_i,),\hfill \end{array}$$
(49)
where $`i=1,\mathrm{},r`$. Recall that the cohomology algebra $`H^{}(\mathrm{BU}k,)`$ is generated freely over $``$ by elements $`\gamma _{\mathrm{U}k}^{(2j)}H^{2j}(\mathrm{BU}k,)`$, $`j=1,\mathrm{},k`$, see . We denote
$$\gamma _{\mathrm{U}k}=1+\gamma _{\mathrm{U}k}^{(2)}+\mathrm{}+\gamma _{\mathrm{U}k}^{(2k)}.$$
(50)
The generators $`\gamma _{\mathrm{U}k_i}^{(2j)}`$ define elements
$`\stackrel{~}{\gamma }_{J,i}^{(2j)}`$ $`=`$ $`\left(\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\gamma _{\mathrm{U}k_i}^{(2j)},`$ (51)
$`\gamma _{J,i}^{(2j)}`$ $`=`$ $`\left(\mathrm{B}j_J\right)^{}\left(\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\gamma _{\mathrm{U}k_i}^{(2j)}`$ (52)
of $`H^{2j}(\mathrm{BU}J,)`$ and $`H^{2j}(\mathrm{BSU}J,)`$, respectively. We denote
$`\stackrel{~}{\gamma }_{J,i}`$ $`=`$ $`1+\stackrel{~}{\gamma }_{J,i}^{(2)}+\mathrm{}+\stackrel{~}{\gamma }_{J,i}^{(2k_i)},i=1,\mathrm{},r,`$ (53)
$`\gamma _{J,i}`$ $`=`$ $`1+\gamma _{J,i}^{(2)}+\mathrm{}+\gamma _{J,i}^{(2k_i)},i=1,\mathrm{},r,`$ (54)
as well as $`\stackrel{~}{\gamma }_J=(\stackrel{~}{\gamma }_{J,1},\mathrm{},\stackrel{~}{\gamma }_{J,r})`$ and $`\gamma _J=(\gamma _{J,1},\mathrm{},\gamma _{J,r})`$.
###### Lemma 5.6
The cohomology algebra $`H^{}(\mathrm{BU}J,)`$ is generated freely over $``$ by the elements $`\stackrel{~}{\gamma }_{J,i}^{(2j)}`$, $`j=1,\mathrm{},k_i`$, $`i=1,\mathrm{},r`$.
Proof: Consider the isomorphism and induced isomorphism
$$\begin{array}{ccccc}\mathrm{U}J& \stackrel{d_r}{}& \mathrm{\Pi }_i\mathrm{U}J& \stackrel{\mathrm{\Pi }_i\mathrm{pr}_i^{\mathrm{U}J}}{}& \mathrm{\Pi }_i\mathrm{U}k_i,\\ H^{}(\mathrm{BU}J,)& \stackrel{d_r^{}}{}& H^{}(\mathrm{\Pi }_i\mathrm{BU}J,)& \stackrel{\left(\mathrm{\Pi }_i\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}}{}& H^{}(\mathrm{\Pi }_i\mathrm{BU}k_i,).\end{array}$$
(55)
where $`d_r`$ denotes $`r`$-fold diagonal embedding and $`\mathrm{\Pi }_i`$ is a shorthand notation for $`\mathrm{\Pi }_{i=1}^r`$. Due to the Kรผnneth Theorem for cohomology \[25, Ch. XIII, Cor. 11.3\], $`H^{}(\mathrm{\Pi }_i\mathrm{BU}k_i,)`$ is generated freely over $``$ by the elements
$$1_{\mathrm{BU}k_1}\times \mathrm{}\times 1_{\mathrm{BU}k_{i1}}\times \gamma _{\mathrm{U}k_i}^{(2j)}\times 1_{\mathrm{BU}k_{i+1}}\times \mathrm{}\times 1_{\mathrm{BU}k_r},j=1,\mathrm{},k_i,i=1,\mathrm{},r.$$
Here $`\times `$ stands for the cohomology cross product, and $`1_{\mathrm{BU}k_i}`$ denotes the generator of $`H^0(\mathrm{BU}k_i,)`$. The assertion follows by applying (55) to these generators and using
$$d_r^{}(\alpha _1\times \mathrm{}\times \alpha _r)=\alpha _1\mathrm{}\alpha _r\alpha _iH^{}(\mathrm{BU}k_i,).\text{ }$$
(56)
###### Lemma 5.7
$`\left(\mathrm{B}j_J\right)^{}`$ is surjective.
Proof: According to (29) and due to $`\mathrm{U}J/\mathrm{SU}J\mathrm{U1}`$, $`\mathrm{B}j_J`$ is the projection in a principal bundle
$$\mathrm{U1}\mathrm{BSU}J\stackrel{\mathrm{B}j_J}{}\mathrm{BU}J.$$
(57)
Denote this bundle by $`\eta `$. Due to $`\pi _1(\mathrm{BU}J)\pi _0(\mathrm{U}J)=0`$, $`\eta `$ is orientable, see \[9, Def. 7.3.3\]. Therefore, it induces a Gysin sequence, see \[9, ยง7.3.1\],
$$\begin{array}{c}H^1(\mathrm{BU}J,)\stackrel{\left(\mathrm{B}j_J\right)^{}}{}H^1(\mathrm{BSU}J,)\stackrel{\sigma ^{}}{}H^0(\mathrm{BU}J,)\stackrel{c_1(\eta )}{}H^2(\mathrm{BU}J,)\hfill \\ \stackrel{\left(\mathrm{B}j_J\right)^{}}{}H^2(\mathrm{BSU}J,)\stackrel{\sigma ^{}}{}H^1(\mathrm{BU}J,)\stackrel{c_1(\eta )}{}H^3(\mathrm{BU}J,)\hfill \\ \stackrel{\left(\mathrm{B}j_J\right)^{}}{}H^3(\mathrm{BSU}J,)\stackrel{\sigma ^{}}{}H^2(\mathrm{BU}J,)\stackrel{c_1(\eta )}{}H^4(\mathrm{BU}J,)\mathrm{}.\hfill \end{array}$$
(58)
(On the level of differential forms, $`\sigma ^{}`$ is given by integration over the fibre.) If $`\eta `$ was trivial, $`\pi _1(\mathrm{BSU}J)`$ would coincide with $`\pi _1(\mathrm{BU}J\times \mathrm{U1})=`$, which would contradict Theorem 5.3. Hence, $`\eta `$ is nontrivial, so that $`c_1(\eta )0`$. Since $`H^{}(\mathrm{BU}J,)`$ has no zero divisors by Lemma 5.6, it follows that multiplication by $`c_1(\eta )`$ is an injective operation on $`H^{}(\mathrm{BU}J,)`$. Then exactness of the Gysin sequence (58) implies that the homomorphism $`\sigma ^{}`$ is trivial and, therefore, $`\left(\mathrm{B}j_J\right)^{}`$ is surjective.
Lemmas 5.6 and 5.7 imply the following corollary.
###### Corollary 5.8
The cohomology algebra $`H^{}(\mathrm{BSU}J,)`$ is generated over $``$ by the elements $`\gamma _{J,i}^{(2j)}`$, $`j=1,\mathrm{},k_i`$, $`i=1,\mathrm{},r`$.
Remark: The generators $`\gamma _{J,i}^{(2)}`$ of $`H^{}(\mathrm{BSU}J,)`$ are subject to a relation which is, however, irrelevant for our purposes. For the sake of completeness, we derive this relation in Appendix B.
### 5.5 Generator for $`H^1(\mathrm{BSU}J,_g)`$
Since $`\mathrm{BSU}J`$ is connected, $`H^1(\mathrm{BSU}J,_g)`$ can be computed by means of the Hurewicz and the Universal Coefficient Theorems:
$$\begin{array}{ccc}\hfill H^1(\mathrm{BSU}J,_g)& & \text{Hom}(H_1(\mathrm{BSU}J),_g)\text{Ext}(H_0(\mathrm{BSU}J),_g)\hfill \\ & & \text{Hom}(\pi _1(\mathrm{BSU}J),_g)\text{Ext}(,_g).\hfill \end{array}$$
Due to Theorem 5.3, $`\pi _1(\mathrm{BSU}J)\pi _0(\mathrm{SU}J)_g`$. Moreover, $`\text{Ext}(,_g)=0`$. Hence, $`H^1(\mathrm{BSU}J,_g)`$ is isomorphic to $`_g`$. It is therefore generated by a single element. Apparently, we are free to choose any of the generators to work with. However, there exists a relation between this generator and the generators $`\gamma _{J,i}^{(2)}`$ of $`H^2(\mathrm{BSU}J,)`$. This can be seen as follows. Consider the short exact sequence
$$0\stackrel{\mu _g}{}\stackrel{\varrho _g}{}_g0,$$
(59)
where $`\mu _g`$ denotes multiplication by $`g`$ and $`\varrho _g`$ reduction modulo $`g`$. This induces a long exact sequence (see \[8, ยงIV.5\])
$$\mathrm{}\stackrel{\beta _g}{}H^i(,)\stackrel{\mu _g}{}H^i(,)\stackrel{\varrho _g}{}H^i(,_g)\stackrel{\beta _g}{}H^{i+1}(,)\stackrel{\mu _g}{}\mathrm{}.$$
(60)
Here we have denoted the coefficient homomorphisms induced by $`\mu _g`$ and $`\varrho _g`$ by the same letters, i.e., $`\mu _g`$ maps $`\alpha g\alpha `$ and $`\varrho _g`$ maps $`\alpha \alpha \text{mod}g`$. Usually, the connecting homomorphism $`\beta _g`$ is called Bockstein homomorphism. Application of $`\beta _g`$ to an arbitrary generator of $`H^1(\mathrm{BSU}J,_g)`$ yields an element of $`H^2(\mathrm{BSU}J,)`$ which can be expressed in terms of the $`\gamma _{J,i}^{(2)}`$. In order to keep track of this relation, we have to choose a specific generator of $`H^1(\mathrm{BSU}J,_g)`$. This will be constructed now.
Consider the homomorphism $`\lambda _J^\mathrm{S}:\mathrm{SU}J_g`$ and the induced homomorphism
$$\left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}:H^1(\mathrm{B}_g,_g)H^1(\mathrm{BSU}J,_g).$$
(61)
Due to Lemma 5.2, $`\lambda _{J}^{\mathrm{S}}{}_{}{}^{}:\pi _0(\mathrm{SU}J)\pi _0(_g)`$ is an isomorphism. Hence, so is $`\left(\mathrm{B}\lambda _J^\mathrm{S}\right)_{}:\pi _1(\mathrm{BSU}J)\pi _1(\mathrm{B}_g)`$. Then the Hurewicz and Universal Coefficient Theorems imply that (61) is an isomorphism. Thus, generators of $`H^1(\mathrm{BSU}J,_g)`$ can be obtained as the images of generators of $`H^1(\mathrm{B}_g,_g)`$ under $`\left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}`$.
###### Lemma 5.9
There exists a unique element $`\delta _gH^1(\mathrm{B}_g,_g)`$ such that
$$\beta _g(\delta _g)=\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}.$$
(62)
This element is a generator of $`H^1(\mathrm{B}_g,_g)`$.
Proof: First we notice that both $`\beta _g(\delta _g)`$ and $`\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$ are elements of $`H^2(\mathrm{B}_g,_g)`$ so that Eq. (62) makes sense.
Next, consider the following portion of the exact sequence (60):
$$\mathrm{}H^1(\mathrm{B}_g,)\stackrel{\varrho _g}{}H^1(\mathrm{B}_g,_g)\stackrel{\beta _g}{}H^2(\mathrm{B}_g,)\stackrel{\mu _g}{}H^2(\mathrm{B}_g,)\mathrm{}.$$
(63)
To determine the cohomology groups, we note that a model for $`\mathrm{B}_g`$ is given by the lens space $`\mathrm{L}_g^{\mathrm{}}`$, see Appendix A. Thus, from Eq. (139) in this appendix we infer
$`H^1(\mathrm{B}_g,)`$ $`=`$ $`0`$ (64)
$`H^2(\mathrm{B}_g,)`$ $`=`$ $`_g.`$ (65)
Due to (64), $`\beta _g`$ is injective in (63). Due to (65), $`\mu _g`$ is trivial in (63) so that $`\beta _g`$ is also surjective there. Hence, we can define $`\delta _g=\beta _g^1\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$. In order to check that this is a generator, we consider $`J^{}=((1),(g))\mathrm{K}(g)`$. We observe that $`_g\mathrm{SU}J^{}`$, $`\mathrm{U1}\mathrm{U}J^{}`$, and that $`j_g`$ corresponds, by virtue of these isomorphisms, to $`j_J^{}:\mathrm{SU}J^{}\mathrm{U}J^{}`$. Then Lemma 5.7 implies that $`\left(\mathrm{B}j_g\right)^{}`$ is surjective. In particular, $`H^2(\mathrm{B}_g,)`$ is generated by $`\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$. Hence, $`H^1(\mathrm{B}_g,_g)`$ is generated by $`\delta _g`$.
We define
$$\delta _J=\left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}\delta _g.$$
(66)
Then Lemma 5.9 yields the following corollary.
###### Corollary 5.10
$`H^1(\mathrm{BSU}J,_g)`$ is generated by $`\delta _J`$, where $`\beta _g(\delta _J)=\left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$.
### 5.6 The Relation between Generators
In this subsection, we are going to derive the relation between $`\delta _J`$ and the $`\gamma _{J,i}^{(2)}`$, i.e., to compute $`\beta _g(\delta _J)`$ in terms of the latter.
For any topological space $`X`$, let $`H_0^{\mathrm{even}}(X,)`$ denote the subset of $`H^{\mathrm{even}}(X,)`$ consisting of elements of the form $`1+\alpha ^{(2)}+\alpha ^{(4)}+\mathrm{}`$ . For any finite sequence of nonnegative integers $`๐=(a_1,\mathrm{},a_s)`$, define a polynomial function
$$E_๐:\underset{i=1}{\overset{s}{}}H_0^{\mathrm{even}}(X,)H_0^{\mathrm{even}}(X,),(\alpha _1,\mathrm{},\alpha _s)\alpha _1^{a_1}\mathrm{}\alpha _s^{a_s},$$
(67)
where powers are taken w.r.t. the cup product. By construction, for any map $`f:XY`$,
$$f^{}E_๐(\alpha _1,\mathrm{},\alpha _s)=E_๐(f^{}\alpha _1,\mathrm{},f^{}\alpha _s).$$
(68)
Let us derive explicit expressions for the components of $`E_๐`$ of $`2`$nd and $`4`$th degree.
###### Lemma 5.11
Let $`\alpha =(\alpha _1,\mathrm{},\alpha _s)_{i=1}^sH_0^{\mathrm{even}}(X,)`$. Then
$`E_๐^{(2)}(\alpha )`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{s}{}}}a_i\alpha _i^{(2)},`$ (69)
$`E_๐^{(4)}(\alpha )`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{s}{}}}a_i\alpha _i^{(4)}+{\displaystyle \underset{i=1}{\overset{s}{}}}{\displaystyle \frac{a_i(a_i1)}{2}}\alpha _i^{(2)}\alpha _i^{(2)}+{\displaystyle \underset{i<j=2}{\overset{s}{}}}a_ia_j\alpha _i^{(2)}\alpha _j^{(2)}.`$ (70)
Proof: This is a straightforward computation which we only indicate here. Let $`\alpha `$ be given. Without loss of generality we may assume that the components of $`\alpha _i`$ of degree higher than $`4`$ vanish. For the cup product of elements of this form one has the following formula:
$$\begin{array}{c}\left(1+\beta ^{(2)}+\beta ^{(4)}\right)\left(1+\gamma ^{(2)}+\gamma ^{(4)}\right)\hfill \\ =1+\left(\beta ^{(2)}+\gamma ^{(2)}\right)+\left(\beta ^{(4)}+\gamma ^{(4)}+\beta ^{(2)}\gamma ^{(2)}\right).\hfill \end{array}$$
We can iterate this to obtain
$$\alpha _i^{a_i}=1+\left(a_i\alpha _i\right)+\left(a_i\alpha _i^{(4)}+\frac{a_i(a_i1)}{2}\alpha _i^{(2)}\alpha _i^{(2)}\right)$$
and then to compute the product of all the factors $`\alpha _i^{a_i}`$. This yields the assertion.
As an immediate consequence of (69), for any $`l`$,
$$E_{l๐}^{(2)}=lE_๐^{(2)}.$$
(71)
###### Lemma 5.12
The following two formulae hold:
$`\left(\mathrm{B}i_J\right)^{}\gamma _{\mathrm{U}n}`$ $`=`$ $`E_๐ฆ\left(\stackrel{~}{\gamma }_J\right),`$ (72)
$`\left(\mathrm{B}\lambda _J\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$ $`=`$ $`E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right).`$ (73)
Proof: First, consider (72). We decompose $`i_J`$ as follows:
$$i_J:\mathrm{U}J\stackrel{d_r}{}\mathrm{\Pi }_i\mathrm{U}J\stackrel{\mathrm{\Pi }_i\mathrm{pr}_i^{\mathrm{U}J}}{}\mathrm{\Pi }_i\mathrm{U}k_i\stackrel{\mathrm{\Pi }_id_{m_i}}{}\mathrm{\Pi }_i(\mathrm{U}k_i\stackrel{m_i}{\times \mathrm{}\times }\mathrm{U}k_i)\stackrel{j}{}\mathrm{U}n.$$
(74)
Here $`d_r`$, $`d_{m_i}`$ denote $`r`$-fold and $`m_i`$-fold diagonal embedding, respectively, and $`j`$ stands for the natural (blockwise) embedding. According to (74), $`\left(\mathrm{B}i_J\right)^{}`$ decomposes as
$$\begin{array}{c}\left(\mathrm{B}i_J\right)^{}:H^{}(\mathrm{BU}n,)\stackrel{\left(\mathrm{B}j\right)^{}}{}H^{}(\mathrm{\Pi }_i(\mathrm{BU}k_i\stackrel{m_i}{\times \mathrm{}\times }\mathrm{BU}k_i),)\hfill \\ \stackrel{\left(\mathrm{\Pi }_id_{m_i}\right)^{}}{}H^{}(\mathrm{\Pi }_i\mathrm{BU}k_i,)\stackrel{\left(\mathrm{\Pi }_i\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}}{}H^{}(\mathrm{\Pi }_i\mathrm{BU}J,)\stackrel{d_r^{}}{}H^{}(\mathrm{BU}J,).\hfill \end{array}$$
(75)
Under the assumption that we have chosen the generators $`\gamma _{\mathrm{U}k}^{(2j)}`$ for different $`k`$ in a consistent way (namely, such that the universal properties of Chern classes hold for the characteristic classes defined by these elements), there holds the relation
$$\left(\mathrm{B}j\right)^{}\gamma _{\mathrm{U}n}=(\gamma _{\mathrm{U}k_1}\stackrel{m_1}{\times \mathrm{}\times }\gamma _{\mathrm{U}k_1})\times \mathrm{}\times (\gamma _{\mathrm{U}k_r}\stackrel{m_r}{\times \mathrm{}\times }\gamma _{\mathrm{U}k_r}).$$
(76)
Using this, as well as (56), we obtain
$$\begin{array}{ccc}\hfill \left(\mathrm{B}i_J\right)^{}\gamma _{\mathrm{U}n}& =& d_r^{}\left(\mathrm{\Pi }_i\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\left(\mathrm{\Pi }_id_{m_i}\right)^{}\left(\mathrm{B}j\right)^{}\gamma _{\mathrm{U}n}\hfill \\ & =& d_r^{}\left(\mathrm{\Pi }_i\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\left(\mathrm{\Pi }_id_{m_i}\right)^{}\left((\gamma _{\mathrm{U}k_1}\stackrel{m_1}{\times \mathrm{}\times }\gamma _{\mathrm{U}k_1})\times \mathrm{}\times (\gamma _{\mathrm{U}k_r}\stackrel{m_r}{\times \mathrm{}\times }\gamma _{\mathrm{U}k_r})\right)\hfill \\ & =& d_r^{}\left(\mathrm{\Pi }_i\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\left(\gamma _{\mathrm{U}k_1}^{m_1}\times \mathrm{}\times \gamma _{\mathrm{U}k_r}^{m_r}\right)\hfill \\ & =& d_r^{}\left(\stackrel{~}{\gamma }_{J,1}^{m_1}\times \mathrm{}\times \stackrel{~}{\gamma }_{J,r}^{m_r}\right)\hfill \\ & =& \stackrel{~}{\gamma }_{J,1}^{m_1}\mathrm{}\stackrel{~}{\gamma }_{J,r}^{m_r}.\hfill \end{array}$$
This yields (72). Now consider (73). The commutative diagram (38) implies
$$\left(\mathrm{B}\lambda _J\right)^{}\left(\mathrm{B}p_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}=\left(\mathrm{B}i_J\right)^{}\left(\mathrm{B}det_{\mathrm{U}n}\right)^{}\gamma _{\mathrm{U1}}^{(2)}.$$
(77)
We have
$`\left(\mathrm{B}det_{\mathrm{U}n}\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$ $`=`$ $`\gamma _{\mathrm{U}n}^{(2)},`$ (78)
$`\left(\mathrm{B}p_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}`$ $`=`$ $`g\gamma _{\mathrm{U1}}^{(2)}.`$ (79)
Formula (79) follows, by virtue of the Hurewicz and the Universal Coefficient Theorems, from the fact that the homomorphism $`\left(\mathrm{B}p_g\right)_{}:\pi _2(\mathrm{BU1})\pi _2(\mathrm{BU1})`$ is given by multiplication by $`g`$. Inserting Eqs. (78) and (79) into (77) we obtain
$$\begin{array}{ccccc}\hfill g\left(\mathrm{B}\lambda _J\right)^{}\gamma _{\mathrm{U1}}^{(2)}& =& E_๐ฆ^{(2)}\left(\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{72}\text{)}\hfill \\ & =& gE_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{71}\text{).}\hfill \end{array}$$
Since this relation holds in $`H^2(\mathrm{BU}J,)`$ which is free Abelian, it implies (73).
###### Theorem 5.13
There holds the relation $`\beta _g(\delta _J)=E_{\stackrel{~}{๐ฆ}}^{(2)}(\gamma _J)`$.
Proof: We compute
$$\begin{array}{ccccc}\hfill \beta _g(\delta _J)& =& \left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & |& \text{by Corollary }\text{5.10}\hfill \\ & =& \left(\mathrm{B}j_J\right)^{}\left(\mathrm{B}\lambda _J\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & |& \text{by (}\text{39}\text{)}\hfill \\ & =& \left(\mathrm{B}j_J\right)^{}E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{73}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\gamma _J\right).\hfill & & \end{array}$$
### 5.7 Characteristic Classes for $`\mathrm{SU}J`$-Bundles
Using the cohomology elements $`\gamma _{J,i}^{(2j)}`$ and $`\delta _J`$ constructed above, we define the following characteristic classes for $`\mathrm{SU}J`$-bundles over a manifold $`M`$:
$`\alpha _{J,i}`$ $`:`$ $`\text{Bun}(M,\mathrm{SU}J)H_0^{\mathrm{even}}(M,),Q\left(f_Q\right)^{}\gamma _{J,i},i=1,\mathrm{},r`$ (80)
$`\xi _J`$ $`:`$ $`\text{Bun}(M,\mathrm{SU}J)H^1(M,_g),Q\left(f_Q\right)^{}\delta _J.`$ (81)
Sorted by degree, $`\alpha _{J,i}(Q)=1+\alpha _{J,i}^{(2)}(Q)+\mathrm{}+\alpha _{J,i}^{(2k_i)}(Q)`$, where $`\alpha _{J,i}^{(2j)}(Q)=\left(f_Q\right)^{}\gamma _{J,i}^{(2j)}`$. Moreover, we introduce the notation $`\alpha _J(Q)=(\alpha _{J,1}(Q),\mathrm{},\alpha _{J,r}(Q))`$. Then
$$\alpha _J(Q)=\left(f_Q\right)^{}\gamma _J$$
(82)
and $`\alpha _J`$ can be viewed as a map from $`\text{Bun}(M,\mathrm{SU}J)`$ to the set
$$H^{(J)}(M,)=\underset{i=1}{\overset{r}{}}\underset{j=1}{\overset{k_i}{}}H^{2j}(M,).$$
(83)
By construction, the relation which holds for $`\gamma _J`$ and $`\delta _J`$ carries over to the characteristic classes $`\alpha _J`$ and $`\xi _J`$. By virtue of (68), from Theorem 5.13 we infer
$$E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\alpha _J(Q)\right)=\beta _g\left(\xi _J(Q)\right)Q\text{Bun}(M,\mathrm{SU}J).$$
(84)
In order to derive expressions for $`\alpha _J`$ and $`\xi _J`$ in terms of the ordinary characteristic classes for $`\mathrm{U}k_i`$-bundles and $`_g`$-bundles, let $`Q\text{Bun}(M,SUJ)`$. There are two kinds of principal bundles associated in a natural way to $`Q`$: The $`\mathrm{U}k_i`$-bundles $`Q^{[\mathrm{pr}_i^{\mathrm{U}J}j_J]}`$, $`i=1,\mathrm{},r`$, and the $`_g`$-bundle $`Q^{[\lambda _J^\mathrm{S}]}`$. For the first ones, using (27) and (52) we compute
$$c\left(Q^{[\mathrm{pr}_i^{\mathrm{U}J}j_J]}\right)=\left(f_{Q^{[\mathrm{pr}_i^{\mathrm{U}J}j_J]}}\right)^{}\gamma _{\mathrm{U}k_i}=\left(f_Q\right)^{}\left(\mathrm{B}j_J\right)^{}\left(\mathrm{Bpr}_i^{\mathrm{U}J}\right)^{}\gamma _{\mathrm{U}k_i}=\left(f_Q\right)^{}\gamma _{J,i},$$
so that
$$\alpha _{J,i}\left(Q\right)=c\left(Q^{[\mathrm{pr}_i^{\mathrm{U}J}j_J]}\right),i=1,\mathrm{},r.$$
(85)
As for the second one, let $`\chi _g`$ denote the characteristic class for $`_g`$-bundles over $`M`$ defined by the generator $`\delta _gH^1(\mathrm{B}_g,_g)`$, i.e.,
$$\chi _g(R)=\left(f_R\right)^{}\delta _g,R\text{Bun}(M,_g).$$
(86)
Then (27) and (66) yield $`\chi _g\left(Q^{[\lambda _J^\mathrm{S}]}\right)=\left(f_{Q^{[\lambda _J^\mathrm{S}]}}\right)^{}\delta _g=\left(f_Q\right)^{}\left(\mathrm{B}\lambda _J^\mathrm{S}\right)^{}\delta _g=\left(f_Q\right)^{}\delta _J`$. Consequently,
$$\xi _J\left(Q\right)=\chi _g\left(Q^{[\lambda _J^\mathrm{S}]}\right).$$
(87)
### 5.8 Classification of $`\mathrm{SU}J`$-Bundles
We denote
$$\mathrm{K}(M)_J=\left\{(\alpha ,\xi )H^{(J)}(M,)\times H^1(M,_g)\right|E_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha )=\beta _g(\xi )\}.$$
(88)
###### Theorem 5.14
Let $`M`$ be a manifold, $`dimM4`$, and let $`J\mathrm{K}(n)`$. Then the characteristic classes $`\alpha _J`$ and $`\xi _J`$ define a bijection from $`\text{Bun}(M,\mathrm{SU}J)`$ onto $`\mathrm{K}(M)_J`$.
Proof: The map is injective by Corollary 5.5. In the following lemma we prove that it is also surjective.
###### Lemma 5.15
Let $`M`$ be a manifold, $`dimM4`$, and let $`J\mathrm{K}(n)`$. Let $`(\alpha ,\xi )\mathrm{K}(M)_J`$. Then there exists $`Q\text{Bun}(M,\mathrm{SU}J)`$ such that $`\alpha _J(Q)=\alpha `$ and $`\xi _J(Q)=\xi `$.
Proof: We give a construction of $`Q`$ in terms of $`\mathrm{U}k_i`$ and $`_g`$-bundles. There exists $`R\text{Bun}(M,_g)`$ such that $`\chi _g(R)=\xi `$. Due to $`dimM4`$, there exist also $`Q_i\text{Bun}(M,\mathrm{U}k_i)`$ such that $`c(Q_i)=\alpha _i`$, $`i=1,\mathrm{},r`$. Define $`\stackrel{~}{Q}=Q_1\times _M\mathrm{}\times _MQ_r`$ (Whitney, or fibre, product). By identifying $`\mathrm{U}k_1\times \mathrm{}\times \mathrm{U}k_r`$ with $`\mathrm{U}J`$, $`\stackrel{~}{Q}`$ becomes a $`\mathrm{U}J`$-bundle. Then
$$\stackrel{~}{Q}^{[\mathrm{pr}_i^{\mathrm{U}J}]}Q_i,i=1,\mathrm{},r.$$
(89)
Consider the $`\mathrm{U1}`$-bundles $`\stackrel{~}{Q}^{[\lambda _J]}`$ and $`R^{[j_g]}`$ associated to $`\stackrel{~}{Q}`$ and $`R`$, respectively. Assume, for a moment, that they are isomorphic. Then $`R`$ is a subbundle of $`\stackrel{~}{Q}^{[\lambda _J]}`$ with structure group $`_g`$. Let $`Q`$ denote the pre-image of $`R`$ under the natural bundle morphism $`\stackrel{~}{Q}\stackrel{~}{Q}^{[\lambda _J]}`$, see (25). This is a subbundle of $`\stackrel{~}{Q}`$ with structure group being the pre-image of $`_g`$ under $`\lambda _J`$, i.e., with structure group $`\mathrm{SU}J`$. Using (85), $`Q^{[j_J]}=\stackrel{~}{Q}`$, and (89), for $`i=1,\mathrm{},r`$,
$$\alpha _{J,i}(Q)=c\left(Q^{[\mathrm{pr}_i^{\mathrm{U}J}j_J]}\right)=c\left(\left(Q^{[j_J]}\right)^{[\mathrm{pr}_i^{\mathrm{U}J}]}\right)=c\left(\stackrel{~}{Q}^{[\mathrm{pr}_i^{\mathrm{U}J}]}\right)=c\left(Q_i\right)=\alpha _i.$$
Moreover, by construction of $`Q`$, $`Q^{[\lambda _J^\mathrm{S}]}R`$. Thus, (87) yields $`\xi _J\left(Q\right)=\chi _J\left(R\right)=\xi `$.
It remains to prove $`\stackrel{~}{Q}^{[\lambda _J]}R^{[j_g]}`$. We compute
$$\begin{array}{ccccc}\hfill c_1\left(\stackrel{~}{Q}^{[\lambda _J]}\right)& =& \left(f_{\stackrel{~}{Q}^{[\lambda _J]}}\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & & \\ & =& \left(f_{\stackrel{~}{Q}}\right)^{}\left(\mathrm{B}\lambda _J\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & |& \text{by (}\text{27}\text{)}\hfill \\ & =& \left(f_{\stackrel{~}{Q}}\right)^{}E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{73}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\left(f_{\stackrel{~}{Q}}\right)^{}\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{68}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}(\left(f_{\stackrel{~}{Q}}\right)^{}\left(\mathrm{Bpr}_1^{\mathrm{U}J}\right)^{}\gamma _{\mathrm{U}k_1},\mathrm{},\left(f_{\stackrel{~}{Q}}\right)^{}\left(\mathrm{Bpr}_r^{\mathrm{U}J}\right)^{}\gamma _{\mathrm{U}k_r})\hfill & |& \text{by (}\text{51}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}(\left(f_{\stackrel{~}{Q}^{[\mathrm{pr}_1^{\mathrm{U}J}]}}\right)^{}\gamma _{\mathrm{U}k_1},\mathrm{},\left(f_{\stackrel{~}{Q}^{[\mathrm{pr}_r^{\mathrm{U}J}]}}\right)^{}\gamma _{\mathrm{U}k_r})\hfill & |& \text{by (}\text{27}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}(\left(f_{Q_1}\right)^{}\gamma _{\mathrm{U}k_1},\mathrm{},\left(f_{Q_r}\right)^{}\gamma _{\mathrm{U}k_r})\hfill & |& \text{by (}\text{89}\text{)}\hfill \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}(c(Q_1),\mathrm{},c(Q_r))\hfill & & \\ & =& E_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha ),\hfill & & \end{array}$$
as well as
$$\begin{array}{ccccc}\hfill c_1\left(R^{[j_g]}\right)& =& \left(f_{R^{[j_g]}}\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & & \\ & =& \left(f_R\right)^{}\left(\mathrm{B}j_g\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & |& \text{by (}\text{27}\text{)}\hfill \\ & =& \left(f_R\right)^{}\beta _g(\delta _g)\hfill & |& \text{by (}\text{62}\text{)}\hfill \\ & =& \beta _g\left(f_R\right)^{}\delta _g\hfill & |& \text{by (}\text{68}\text{)}\hfill \\ & =& \beta _g(\chi _g(R))\hfill & |& \text{by (}\text{86}\text{)}\hfill \\ & =& \beta _g(\xi ).\hfill & & \end{array}$$
Thus, due to $`(\alpha ,\xi )\mathrm{K}(M)_J`$, $`c_1(\stackrel{~}{Q}^{[\lambda _J]})=c_1(R^{[j_g]})`$. It follows that, indeed, $`\stackrel{~}{Q}^{[\lambda _J]}R^{[j_g]}`$. This proves the lemma and, therefore, the theorem.
### 5.9 Classification of $`\mathrm{SU}J`$-Subbundles of $`\mathrm{SU}n`$-Bundles
Let $`P`$ be a principal $`\mathrm{SU}n`$-bundle over a manifold $`M`$ and let $`J\mathrm{K}(n)`$. We are going to characterize the subset $`\text{Red}(P,\mathrm{SU}J)\text{Bun}(M,\mathrm{SU}J)`$ in terms of the characteristic classes $`\alpha _J`$ and $`\xi _J`$. Recall that for $`Q\text{Bun}(M,\mathrm{SU}J)`$, $`Q^{[\mathrm{SU}n]}`$ denotes the extension of $`Q`$ by $`\mathrm{SU}n`$.
###### Lemma 5.16
For any $`Q\text{Bun}(M,\mathrm{SU}J)`$, $`c\left(Q^{[\mathrm{SU}n]}\right)=E_๐ฆ\left(\alpha _J(Q)\right)`$.
Proof: Note that $`c\left(Q^{[\mathrm{SU}n]}\right)=c\left(Q^{[\mathrm{U}n]}\right)=c\left(Q^{[i_Jj_J]}\right)`$. Hence,
$$\begin{array}{ccccc}\hfill c\left(Q^{[\mathrm{SU}n]}\right)& =& \left(f_Q\right)^{}\left(\mathrm{B}j_J\right)^{}\left(\mathrm{B}i_J\right)^{}\gamma _{\mathrm{U}n}\hfill & |& \text{by (}\text{27}\text{)}\hfill \\ & =& \left(f_Q\right)^{}\left(\mathrm{B}j_J\right)^{}E_๐ฆ\left(\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{72}\text{)}\hfill \\ & =& E_๐ฆ\left(\left(f_Q\right)^{}\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\gamma }_J\right)\hfill & |& \text{by (}\text{68}\text{)}\hfill \\ & =& E_๐ฆ\left(\left(f_Q\right)^{}\gamma _J\right)\hfill & |& \text{by (}\text{52}\text{)}\hfill \\ & =& E_๐ฆ\left(\alpha _J(Q)\right)\hfill & |& \text{by (}\text{82}\text{).}\hfill \end{array}$$
We define
$$\mathrm{K}(P)_J=\left\{(\alpha ,\xi )\mathrm{K}(M)_J\right|E_๐ฆ(\alpha )=c(P)\}.$$
(90)
###### Theorem 5.17
Let $`P`$ be a principal $`\mathrm{SU}n`$-bundle over a manifold $`M`$, $`dimM4`$, and let $`J\mathrm{K}(n)`$. Then the characteristic classes $`\alpha _J`$, $`\xi _J`$ define a bijection from $`\text{Red}(P,\mathrm{SU}J)`$ onto $`\mathrm{K}(P)_J`$.
Proof: Let $`Q\text{Bun}(M,\mathrm{SU}J)`$. Then $`(\alpha _J(Q),\xi _J(Q))\mathrm{K}(M)_J`$. Lemma 5.16 implies that $`(\alpha _J(Q),\xi _J(Q))\mathrm{K}(P)_J`$ if and only if $`c\left(Q^{[\mathrm{SU}n]}\right)=c(P)`$. Due to $`dimM4`$, the latter is equivalent to $`Q^{[\mathrm{SU}n]}P`$, i.e., to $`Q\text{Red}(P,\mathrm{SU}J)`$.
The equation $`E_๐ฆ(\alpha )=c(P)`$ actually contains the two equations $`E_๐ฆ^{(2)}(\alpha )=0`$ and $`E_๐ฆ^{(4)}(\alpha )=c_2(P)`$. However, under the assumption that $`(\alpha ,\xi )\mathrm{K}(M)_J`$, the first one is redundant, because then, due to (71), $`E_๐ฆ^{(2)}(\alpha )=gE_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha )=g\beta _g(\xi )=0`$. Thus, the relevant equations are
$`E_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha )`$ $`=`$ $`\beta _g(\xi ),`$ (91)
$`E_๐ฆ^{(4)}(\alpha )`$ $`=`$ $`c_2(P),`$ (92)
where $`\alpha H^{(J)}(M,)`$, $`\xi H^1(M,_g)`$. The set of solutions of Eq. (91) yields $`\mathrm{K}(M)_J`$, hence $`\text{Bun}(M,\mathrm{SU}J)`$. The set of solutions of both Eqs. (91) and (92) yields $`\mathrm{K}(P)_J`$ and, therefore, $`\text{Red}(P,\mathrm{SU}J)`$.
This concludes the classification of Howe subbundles of $`P`$, i.e., Step 2 of our programme.
### 5.10 Examples
We are going to determine $`\mathrm{K}(P)_J`$, i.e., to solve the system of equations (91) and (92), for several choices of $`J`$ and for base manifolds $`M=\mathrm{S}^4,\mathrm{S}^2\times \mathrm{S}^2,\mathrm{T}^4`$, and $`\mathrm{L}_p^3\times \mathrm{S}^1`$. Here $`\mathrm{L}_p^3`$ denotes the $`3`$-dimensional lens space which is defined to be the quotient of the restriction of the natural action of $`\mathrm{U1}`$ on the sphere $`\mathrm{S}^3^2`$ to the subgroup $`_p`$. Note that $`\mathrm{L}_p^3`$ is orientable. We remark that there are more general lens spaces, even in $`3`$ dimensions.
#### Preliminary Remarks
Due to compactness and orientability, $`H^4(M,)`$. Let us derive the Bockstein homomorphism $`\beta _g:H^1(M,_g)H^2(M,)`$. Since for products of spheres the integer-valued cohomology is torsion-free, $`\beta _g`$ is trivial here. On the other hand, consider $`M=\mathrm{L}_p^3\times \mathrm{S}^1`$. From the exact homotopy sequence induced by the fibration $`_p\mathrm{S}^3\mathrm{L}_p^3`$ we infer
$$\pi _1\left(\mathrm{L}_p^3\right)_p.$$
(93)
According to the Hurewicz and Universal Coefficient Theorems, then
$$H^1(\mathrm{L}_p^3,_g)\text{Hom}(\pi _1\left(\mathrm{L}_p^3\right),_g)\text{Hom}(_p,_g)_{p,g},$$
(94)
where $`p,g`$ denotes the greatest common divisor of $`p`$ and $`g`$. Let $`\gamma _{\mathrm{L}_p^3;_g}^{(1)}`$ and $`\gamma _{\mathrm{S}^1}^{(1)}`$ be generators of $`H^1(\mathrm{L}_p^3,_g)`$ and $`H^1(\mathrm{S}^1,)`$, respectively. Due to the Kรผnneth Theorem for cohomology \[25, Ch. XIII, Thm. 11.2\] and (94),
$$\begin{array}{ccc}\hfill H^1(\mathrm{L}_p^3\times \mathrm{S}^1,_g)& & H^1(\mathrm{L}_p^3,_g)H^0(\mathrm{S}^1,)H^0(\mathrm{L}_p^3,_g)H^1(\mathrm{S}^1,)\hfill \\ & & _{p,g}_g,\hfill \end{array}$$
(95)
where the first factor is generated by $`\gamma _{\mathrm{L}_p^3;_g}^{(1)}\times 1_{\mathrm{S}^1}`$ and the second one by $`1_{\mathrm{L}_p^3;_g}\times \gamma _{\mathrm{S}^1}^{(1)}`$. Consider the following portion of the exact sequence (60):
$$H^1(\mathrm{L}_p^3,)\stackrel{\varrho _g}{}H^1(\mathrm{L}_p^3,_g)\stackrel{\beta _g}{}H^2(\mathrm{L}_p^3,).$$
(96)
One has, see \[8, ยง10, p. 363\],
$$H^1(\mathrm{L}_p^3,)=0,H^2(\mathrm{L}_p^3,)_p$$
(97)
(the first equality follows also from (93)). In view of (97) and (94), (96) implies
$$\beta _g\left(\gamma _{\mathrm{L}_p^3;_g}^{(1)}\right)=\frac{p}{p,g}\gamma _{\mathrm{L}_p^3;}^{(2)},$$
(98)
where $`\gamma _{\mathrm{L}_p^3;}^{(2)}`$ is an appropriately chosen generator of $`H^2(\mathrm{L}_p^3,)`$. Moreover, due to (97), $`\beta _g(1_{\mathrm{L}_p^3;_g})=0`$. Thus, we obtain
$$\beta _g\left(\gamma _{\mathrm{L}_p^3;_g}^{(1)}\times 1_{\mathrm{S}^1}\right)=\frac{p}{p,g}\gamma _{\mathrm{L}_p^3;}^{(2)}\times 1_{\mathrm{S}^1},\beta _g\left(1_{\mathrm{L}_p^3;_g}\times \gamma _{\mathrm{S}^1}^{(1)}\right)=0.$$
(99)
Finally, note that via the Kรผnneth Theorem, (97) implies $`H^2(\mathrm{L}_p^3\times \mathrm{S}^1,)_p`$. Since $`H^4(\mathrm{L}_p^3\times \mathrm{S}^1,)`$ is torsion-free, then the cup product $`\alpha _1^{(2)}\alpha _2^{(2)}`$ is trivial.
Now let us discuss some special choices for $`J`$. For brevity, we write $`J`$ in the form $`J=(k_1,\mathrm{},k_r|m_1,\mathrm{},m_r)`$.
#### $`๐ฑ\mathbf{=}\mathbf{(}\mathrm{๐}\mathbf{|}๐\mathbf{)}\mathbf{}๐\mathbf{(}๐\mathbf{)}`$
Here $`\mathrm{SU}J=_n`$, the center of $`\mathrm{SU}n`$. Moreover, $`g=n`$. Variables are $`\xi H^1(M,_n)`$ and $`\alpha =1+\alpha ^{(2)}`$, $`\alpha ^{(2)}H^2(M,)`$. The system of equations (91) and (92) reads
$`\alpha ^{(2)}`$ $`=`$ $`\beta _n(\xi )`$ (100)
$`{\displaystyle \frac{n(n1)}{2}}\alpha ^{(2)}\alpha ^{(2)}`$ $`=`$ $`c_2(P).`$ (101)
Note that we have used Lemma 5.11. Eq. (100) merely expresses $`\alpha ^{(2)}`$ in terms of $`\xi `$. In particular, it yields $`n\alpha ^{(2)}=0`$, so that Eq. (101) requires $`c_2(P)=0`$. As a result, $`\mathrm{K}(P)_J`$ is nonempty iff $`P`$ is trivial and is then parametrized by $`\xi `$. This coincides with what is known about $`_n`$-subbundles of $`\mathrm{SU}n`$-bundles.
#### $`๐ฑ\mathbf{=}\mathbf{(}๐\mathbf{|}\mathrm{๐}\mathbf{)}\mathbf{}๐\mathbf{(}๐\mathbf{)}`$
Here $`\mathrm{SU}J=\mathrm{SU}n`$, the whole group. Due to $`g=1`$, the only variable is $`\alpha =1+\alpha ^{(2)}+\alpha ^{(4)}`$, where $`\alpha ^{(2j)}H^{2j}(M,)`$, $`j=1,2`$. The system of equations (91) and (92) is
$$\alpha ^{(2)}=0,\alpha ^{(4)}=c_2(P).$$
Thus, $`\mathrm{K}(P)_J`$ consists of $`P`$ itself.
#### $`๐ฑ\mathbf{=}\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{|}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{}๐\mathbf{(}\mathrm{๐}\mathbf{)}`$
One can check that $`\mathrm{SU}J`$ has connected components
$$\{\text{diag}(z,z,z^1,z^1)|z\mathrm{U1}\},\{\text{diag}(z,z,z^1,z^1)|z\mathrm{U1}\}.$$
It is, therefore, isomorphic to $`\mathrm{U1}\times _2`$. Variables are $`\xi H^1(M,_2)`$ and $`\alpha _i=1+\alpha _i^{(2)}`$, $`\alpha _i^{(2)}H^2(M,)`$, $`i=1,2`$. The system of equations under consideration is
$`\alpha _1^{(2)}+\alpha _2^{(2)}`$ $`=`$ $`\beta _2(\xi )`$ (102)
$`\alpha _1^{(2)}\alpha _1^{(2)}+\alpha _2^{(2)}\alpha _2^{(2)}+4\alpha _1^{(2)}\alpha _2^{(2)}`$ $`=`$ $`c_2(P).`$ (103)
We solve Eq. (102) w.r.t. $`\alpha _2^{(2)}`$ and insert it into Eq. (103). Since $`H^4(M,)`$ is torsion-free, products including $`\beta _2(\xi )`$ vanish. Thus, we obtain
$$2\alpha _1^{(2)}\alpha _1^{(2)}=c_2(P).$$
(104)
For base manifold $`M=\mathrm{S}^4`$, $`H^2(M,)=0`$. Hence, $`\mathrm{K}(P)_J`$ is nonempty iff $`c_2(P)=0`$, in which case it contains the (necessarily trivial) $`\mathrm{U1}\times _2`$-bundle over $`\mathrm{S}^4`$.
For $`M=\mathrm{L}_p^3\times \mathrm{S}^1`$, in case $`c_2(P)=0`$, $`\mathrm{K}(P)_J`$ is parametrized by $`\xi H^1(M,_g)_{2,p}_2`$ and $`\alpha _1^{(2)}H^2(M,)_p`$. Otherwise it is empty.
For $`M=\mathrm{S}^2\times \mathrm{S}^2`$, $`H^1(M,_g)=0`$. Let $`\gamma _{\mathrm{S}^2}^{(2)}`$ be a generator of $`H^2(\mathrm{S}^2,)`$. Due to the Kรผnneth Theorem, $`H^2(M,)`$, where it is generated by $`\gamma _{\mathrm{S}^2}^{(2)}\times 1_{\mathrm{S}^2}`$ and $`1_{\mathrm{S}^2}\times \gamma _{\mathrm{S}^2}^{(2)}`$. Moreover, $`H^4(M,)`$ is generated by $`\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}`$. Writing
$$\alpha _1^{(2)}=a\gamma _{\mathrm{S}^2}^{(2)}\times 1_{\mathrm{S}^2}+b1_{\mathrm{S}^2}\times \gamma _{\mathrm{S}^2}^{(2)}$$
(105)
with $`a,b`$, Eq. (104) becomes
$$4ab\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}=c_2(P).$$
If $`c_2(P)=0`$, there are two series of solutions: $`a=0`$ and $`b`$ as well as $`a`$ and $`b=0`$. Here $`\mathrm{K}(P)_J`$ is infinite. If $`c_2(P)=4l\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}`$, $`l0`$, then $`a=q`$ and $`b=l/q`$, where $`q`$ runs through the (positive and negative) divisors of $`l`$. Hence, in this case, the cardinality of $`\mathrm{K}(P)_J`$ is twice the number of divisors of $`l`$. If $`c_2(P)`$ is not divisible by $`4`$ then $`\mathrm{K}(P)_J`$ is empty.
Finally, for $`M=\mathrm{T}^4`$ one has $`H^1(M,)_2^4`$ and $`H^2(M,)^6`$. Moreover, $`H^2(M,)`$ is generated by elements $`\gamma _{\mathrm{T}^4;ij}^{(2)}`$, $`1i<j4`$, where
$$\gamma _{\mathrm{T}^4;12}^{(2)}=\gamma _{\mathrm{S}^1}^{(1)}\times \gamma _{\mathrm{S}^1}^{(1)}\times 1_{\mathrm{S}^1}\times 1_{\mathrm{S}^1},\gamma _{\mathrm{T}^4;13}^{(2)}=\gamma _{\mathrm{S}^1}^{(1)}\times 1_{\mathrm{S}^1}\times \gamma _{\mathrm{S}^1}^{(1)}\times 1_{\mathrm{S}^1},\text{etc.,}$$
whereas $`H^4(M,)`$ is generated by $`\gamma _{\mathrm{T}^4}^{(4)}=\gamma _{\mathrm{S}^1}^{(1)}\times \gamma _{\mathrm{S}^1}^{(1)}\times \gamma _{\mathrm{S}^1}^{(1)}\times \gamma _{\mathrm{S}^1}^{(1)}`$. One can check
$$\gamma _{\mathrm{T}^4;ij}^{(2)}\gamma _{\mathrm{T}^4;kl}^{(2)}=ฯต_{ijkl}\gamma _{\mathrm{T}^4}^{(4)},$$
(106)
where $`ฯต_{ijkl}`$ denotes the totally antisymmetric tensor in $`4`$ dimensions. Writing
$$\alpha _1^{(2)}=\underset{1i<j4}{}a_{ij}\gamma _{\mathrm{T}^4;ij}^{(2)}$$
(107)
and using (106), Eq. (104) yields
$$4\left(a_{12}a_{34}a_{13}a_{24}+a_{14}a_{23}\right)\gamma _{\mathrm{T}^4}^{(4)}=c_2(P).$$
Hence, we find that $`\mathrm{K}(P)_J`$ is again nonempty iff $`c_2(P)`$ is divisible by $`4`$, in which case it now has always infinitely many elements.
To conclude, let us point out that, when passing to $`\widehat{\mathrm{K}}(P)`$, one has to identify the symbol containing $`(\alpha _1,\alpha _2)`$ with that containing $`(\alpha _2,\alpha _1)`$.
#### $`๐ฑ\mathbf{=}\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{|}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{}๐\mathbf{(}\mathrm{๐}\mathbf{)}`$
The subgroup $`\mathrm{SU}J`$ of $`\mathrm{SU5}`$ consists of matrices of the form $`\text{diag}(z_1,z_1,z_2,z_2,z_2)`$, where $`z_1,z_2\mathrm{U1}`$ such that $`z_1^2z_2^3=1`$. We can parametrize $`z_1=z^3`$, $`z_2=z^2`$, $`z\mathrm{U1}`$. Hence, $`\mathrm{SU}J`$ is isomorphic to $`\mathrm{U1}`$. Variables are $`\alpha _i=1+\alpha _i^{(2)}`$, $`i=1,2`$. The equations to be solved read
$`2\alpha _1^{(2)}+3\alpha _2^{(2)}`$ $`=`$ $`0`$ (108)
$`\alpha _1^{(2)}\alpha _1^{(2)}+3\alpha _2^{(2)}\alpha _2^{(2)}+6\alpha _1^{(2)}\alpha _2^{(2)}`$ $`=`$ $`c_2(P)`$ (109)
Eq. (108) can be parametrized by $`\alpha _1^{(2)}=3\eta `$, $`\alpha _2^{(2)}=2\eta `$, where $`\eta H^2(M,)`$. Then Eq. (109) becomes
$$15\eta \eta =c_2(P).$$
The discussion of this equation is analogous to that of Eq. (104) above.
#### $`๐ฑ\mathbf{=}\mathbf{(}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{|}\mathrm{๐}\mathbf{,}\mathrm{๐}\mathbf{)}\mathbf{}๐\mathbf{(}\mathrm{๐}\mathbf{)}`$
Here $`\mathrm{SU}J\mathrm{S}(\mathrm{U2}\times \mathrm{U3})`$, the symmetry group of the standard model. In the grand unified $`\mathrm{SU5}`$-model this is the subgroup to which $`\mathrm{SU5}`$ is broken by the heavy Higgs field. Moreover, the subgroup $`\mathrm{SU}J`$ is the centralizer of the subgroup $`\mathrm{SU}(1,1|2,3)`$ discussed above. Variables are $`\alpha _i=1+\alpha _i^{(2)}+\alpha _i^{(4)}`$, where $`\alpha _i^{(2j)}H^{2j}(M,)`$, $`i,j=1,2`$. Eqs. (91) and (92) read
$`\alpha _1^{(2)}+\alpha _2^{(2)}`$ $`=`$ $`0,`$ (110)
$`\alpha _1^{(4)}+\alpha _2^{(4)}+\alpha _1^{(2)}\alpha _2^{(2)}`$ $`=`$ $`c_2(P).`$ (111)
Using (110) to replace $`\alpha _2^{(2)}`$ in (111) we obtain for the latter
$$\alpha _2^{(4)}=c_2(P)\alpha _1^{(4)}+\alpha _1^{(2)}\alpha _1^{(2)}.$$
Thus, $`\mathrm{K}(P)_J`$ can be parametrized by $`\alpha _1`$ (or $`\alpha _2`$), i.e., by the Chern class of one of the factors $`\mathrm{U2}`$ or $`\mathrm{U3}`$. This has been known for a long time .
#### $`๐ฑ\mathbf{=}\mathbf{(}\mathrm{๐}\mathbf{|}\mathrm{๐}\mathbf{)}`$
The subgroup $`\mathrm{SU}J`$ of $`\mathrm{SU4}`$ consists of matrices $`DD`$, where $`D\mathrm{U2}`$ such that $`(detD)^2=1`$. Hence, it has connected components $`\{DD|D\mathrm{SU2}\}`$ and $`\{(iD)(iD)|D\mathrm{SU2}\}`$. One can check that the map $`\mathrm{SU2}\times _4\mathrm{SU}J`$, $`(D,a)e^{2\pi ia/4}D`$, induces an isomorphism from $`(\mathrm{SU2}\times _4)/_2`$ onto $`\mathrm{SU}J`$.
Now consider $`\mathrm{K}(P)_J`$. Variables are $`\xi H^1(M,_2)`$ and $`\alpha =1+\alpha ^{(2)}+\alpha ^{(4)}`$. We have
$`\alpha ^{(2)}`$ $`=`$ $`\beta _2(\xi )`$ (112)
$`\alpha ^{(2)}\alpha ^{(2)}+2\alpha ^{(4)}`$ $`=`$ $`c_2(P)`$ (113)
Eq. (112) fixes $`\alpha ^{(2)}`$ in terms of $`\xi `$. For example, in case $`M=\mathrm{L}_p^3\times \mathrm{S}^1`$, by expanding $`\xi =\xi _\mathrm{L}\gamma _{\mathrm{L}_p^3;_g}^{(1)}\times 1_{\mathrm{S}^1}+\xi _\mathrm{S}1_{\mathrm{L}_p^3}\times \gamma _{\mathrm{S}^1}^{(1)}`$, Eqs. (99) and (112) imply
$$\alpha ^{(2)}=\{\begin{array}{ccc}q\xi _L\gamma _{\mathrm{L}_p^3;}^{(2)}\times 1_{\mathrm{S}^1}& |& p=2q\hfill \\ 0& |& p=2q+1.\hfill \end{array}$$
For general $`M`$, due to (112), Eq. (113) becomes $`2\alpha ^{(4)}=c_2(P)`$. Thus, $`\mathrm{K}(P)_J`$ is nonempty iff $`c_2(P)`$ is even and is then parametrized by $`\xi H^1(M,_2)`$.
Let us point out the following. Consider Eq. (112) which defines the set $`\mathrm{K}(M)_J`$ classifying $`\mathrm{SU}J`$-bundles over $`M`$. In case $`M`$ is a product of spheres (or, more generally, has trivial $`\beta _2`$ in degree $`1`$), this classification coincides with that of $`(\mathrm{SU2}\times _2)`$-bundles, although this group is only locally isomorphic to $`\mathrm{SU}J`$. In case $`M=\mathrm{L}_p^3\times \mathrm{S}^1`$, there is a difference though. Here the $`\mathrm{SU}J`$-bundles which are nontrivial over the factor $`\mathrm{L}_p^3`$ have a magnetic charge, whereas an $`(\mathrm{SU2}\times _2)`$-bundle can never have one.
## 6 Holonomy-Induced Howe Subbundles
In the next step of our programme we have to specify the Howe subbundles which are holonomy-induced. However, in fact, this turns out not to be necessary here because for Howe subbundles of principal $`\mathrm{SU}n`$-bundles the condition of being holonomy-induced is redundant. This will be proved now.
###### Lemma 6.1
Let $`HH^{}\mathrm{SU}n`$ be Howe subgroups. If $`dimH=dimH^{}`$ then $`H=H^{}`$.
Proof: There exist $`J,J^{}\mathrm{K}(n)`$ such that $`H`$ and $`H^{}`$ are conjugate to $`\mathrm{SU}J`$ and $`\mathrm{SU}J^{}`$, respectively. Consider $`\mathrm{U}J`$ and $`\mathrm{U}J^{}`$. Due to $`HH^{}`$, there exists $`D\mathrm{SU}n`$ such that $`D^1\mathrm{U}JD\mathrm{U}J^{}`$. Moreover,
$$dim\mathrm{U}J^{}=dim\mathrm{SU}J^{}+1=dimH^{}+1=dimH+1=dim\mathrm{SU}J+1=dim\mathrm{U}J.$$
(114)
$`\mathrm{U}J^{}`$ being connected and $`D^1\mathrm{U}JD`$ being closed in $`\mathrm{U}J^{}`$, (114) implies $`D^1\mathrm{U}JD=\mathrm{U}J^{}`$. Then $`D^1\mathrm{SU}JD=D^1\left(\mathrm{U}J\mathrm{SU}n\right)D=\left(D^1\mathrm{U}JD\right)\mathrm{SU}n=\mathrm{U}J^{}\mathrm{SU}n=\mathrm{SU}J^{}`$. It follows $`H=H^{}`$.
###### Theorem 6.2
Any Howe subbundle of a principal $`\mathrm{SU}n`$-bundle is holonomy-induced.
Proof: Let $`P`$ be a principal $`\mathrm{SU}n`$-bundle and let $`Q`$ be a Howe subbundle of $`P`$ with structure group $`H`$. Denote the structure group of a connected component of $`Q`$ by $`\stackrel{~}{H}`$. It is easily seen that $`Q`$ is holonomy-induced provided
$$\mathrm{C}_{\mathrm{SU}n}^2(\stackrel{~}{H})=H.$$
(115)
Since $`H`$ is a Howe subgroup, we have $`\stackrel{~}{H}\mathrm{C}_{\mathrm{SU}n}^2(\stackrel{~}{H})H`$. Since $`\stackrel{~}{H}`$ and $`H`$ are of the same dimension, so are $`\mathrm{C}_{\mathrm{SU}n}^2(\stackrel{~}{H})`$ and $`H`$. By virtue of Lemma 6.1, this implies (115).
For the reader who wonders whether there exist Howe subbundles which are not holonomy-induced we give an example here. Consider the Lie group $`\mathrm{SO3}`$. One checks that the subgroup $`H=\{\mathrm{๐}_3,\text{diag}(1,1,1)\}`$ is Howe. Thus, the subbundle $`Q=M\times H`$ of the bundle $`M\times \mathrm{SO3}`$ is Howe. Any connected subbundle $`\stackrel{~}{Q}`$ of $`Q`$ has the center $`\{\mathrm{๐}_3\}`$ as its structure group. Since the center is Howe itself, $`\stackrel{~}{Q}\mathrm{C}_G^2(\{\mathrm{๐}_3\})=\stackrel{~}{Q}Q`$.
## 7 Factorization by $`\mathrm{SU}n`$-Action
In Step 4 of our programme to determine $`\text{Howe}_{}(P)`$, we actually have to take the disjoint union of $`\text{Red}(P,H)`$ over all Howe subgroups $`H`$ of $`\mathrm{SU}n`$ and to factorize by the action of $`\mathrm{SU}n`$. Since $`\mathrm{SU}n`$-action on subbundles conjugates their structure groups, however, it suffices to take the union only over $`\mathrm{SU}J`$, $`J\mathrm{K}(n)`$:
$$\underset{J\mathrm{K}(n)}{}\text{Red}(P,\mathrm{SU}J).$$
(116)
According to this, define
$$\mathrm{K}(P)=\underset{J\mathrm{K}(n)}{}\mathrm{K}(P)_J.$$
(117)
We shall denote the elements of $`\mathrm{K}(P)`$ by $`L`$ and write them in the form $`L=(J;\alpha ,\xi )`$, where $`J\mathrm{K}(n)`$ and $`(\alpha ,\xi )\mathrm{K}(P)_J`$. Due to Theorem 5.17, the collection of characteristic classes $`\{(\alpha _J,\xi _J)|J\mathrm{K}(n)\}`$ defines a bijection from (116) onto $`\mathrm{K}(P)`$. Now we reverse this bijection: Let $`L\mathrm{K}(P)`$, $`L=(J;\alpha ,\xi )`$. Then define $`Q_L`$ to be the isomorphism class of $`\mathrm{SU}J`$-subbundles of $`P`$ which obey
$$\alpha _J(Q_L)=\alpha ,\xi _J(Q_L)=\xi .$$
(118)
###### Lemma 7.1
Let $`L,L^{}\mathrm{K}(P)`$, where $`L=(J;\alpha ,\xi )`$ and $`L^{}=(J^{};\alpha ^{},\xi ^{})`$. There exists $`D\mathrm{SU}n`$ such that $`Q_L^{}=Q_LD`$ (up to isomorphy) if and only if $`\xi ^{}=\xi `$, as well as $`J^{}=\sigma J`$ and $`\alpha ^{}=\sigma \alpha `$ for some permutation $`\sigma `$ of $`1,\mathrm{},r`$.
Proof: For $`J_1,J_2\mathrm{K}(n)`$ and $`D\mathrm{SU}n`$ such that $`D^1\mathrm{SU}J_1D\mathrm{SU}J_2`$, define embeddings
$`h_{J_1J_2}^D`$ $`:`$ $`\mathrm{U}J_1\mathrm{U}J_2,CD^1CD`$
$`h_{J_1J_2}^{D;\mathrm{S}}`$ $`:`$ $`\mathrm{SU}J_1\mathrm{SU}J_2,CD^1CD.`$
First, assume that we are given $`D\mathrm{SU}n`$ such that $`Q_L^{}Q_LD`$. Then $`D^1\mathrm{SU}JD=\mathrm{SU}J^{}`$, so that we can consider the isomorphisms $`h_{JJ^{}}^{D;\mathrm{S}}:\mathrm{SU}J\mathrm{SU}J^{}`$ and $`h_{JJ^{}}^D:\mathrm{U}J\mathrm{U}J^{}`$. One can check that
$$Q_LDQ_L^{[h_{JJ^{}}^{D;\mathrm{S}}]}.$$
(119)
Moreover, there exists a permutation $`\sigma `$ of $`1,\mathrm{},r`$ such that $`h_{JJ^{}}^D`$ maps the $`\sigma (i)`$-th factor of $`\mathrm{U}J`$ isomorphically onto the $`i`$-th factor of $`\mathrm{U}J^{}`$. Then, in particular, $`J^{}=\sigma J`$. Moreover, there exists $`C\mathrm{SU}n`$ such that $`\mathrm{pr}_i^{\mathrm{U}J^{}}h_{JJ^{}}^C=\mathrm{pr}_{\sigma (i)}^{\mathrm{U}J}`$ $`i`$ (in fact, $`C`$ has been constructed in the proof of Lemma 4.2). Then
$$\mathrm{pr}_i^{\mathrm{U}J^{}}j_Jh_{JJ^{}}^{C;\mathrm{S}}=\mathrm{pr}_{\sigma (i)}^{\mathrm{U}J}j_J.$$
(120)
Next, define $`B=DC^1\mathrm{SU}n`$. The corresponding homomorphism $`h_{JJ}^B`$ is an automorphism of $`\mathrm{U}J`$ which leaves each factor invariant separately. One can check that it is an inner automorphism of $`\mathrm{U}J`$. Then $`h_{JJ}^{B;\mathrm{S}}`$ is an inner automorphism of $`\mathrm{SU}J`$ and can, therefore, be generated by an element of the connected component of the identity. Then $`\mathrm{B}h_{JJ}^{B;\mathrm{S}}`$ is homotopic to $`\mathrm{Bid}_{\mathrm{SU}J}\mathrm{id}_{\mathrm{BSU}J}`$. As a consequence, the classifying map of the subbundle $`Q_LB`$, which is, due to (119) and (27), given by $`f_{Q_LB}=f_{Q_L^{[h_{JJ}^{B;\mathrm{S}}]}}=\mathrm{B}h_{JJ}^{B;\mathrm{S}}f_{Q_L}`$, is homotopic to that of $`Q_L`$. It follows $`Q_L^{}Q_LC`$. Then, due to (119),
$$Q_L^{}Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}.$$
(121)
We compute the characteristic classes of $`Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}`$ in terms of those of $`Q_L`$:
$$\begin{array}{ccccc}\hfill \alpha _{J^{},i}\left(Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}\right)& =& c\left(\left(Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}\right)^{[\mathrm{pr}_i^{\mathrm{U}J^{}}j_J]}\right)\hfill & |& \text{by (}\text{85}\text{)}\hfill \\ & =& c\left(Q_L^{[\mathrm{pr}_i^{\mathrm{U}J^{}}j_Jh_{JJ^{}}^{C;\mathrm{S}}]}\right)\hfill & & \\ & =& c\left(Q_L^{[\mathrm{pr}_{\sigma (i)}^{\mathrm{U}J}j_J]}\right)\hfill & |& \text{by (}\text{120}\text{)}\hfill \\ & =& \alpha _{J,\sigma (i)}\left(Q_L\right),\hfill & & \end{array}$$
(122)
as well as, using $`\lambda _J^{}^\mathrm{S}h_{JJ^{}}^{C;\mathrm{S}}=\lambda _J^\mathrm{S}`$ and (87),
$$\xi _J^{}\left(Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}\right)=\chi _g\left(\left(Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}\right)^{[\lambda _J^{}^\mathrm{S}]}\right)=\chi _g\left(Q_L^{[\lambda _J^\mathrm{S}]}\right)=\xi _J\left(Q_L\right).$$
(123)
Thus,
$$\begin{array}{ccccc}\hfill \alpha _i^{}& =& \alpha _{J^{},i}\left(Q_L^{}\right)\hfill & |& \text{by (}\text{118}\text{)}\hfill \\ & =& \alpha _{J^{},i}\left(Q_L^{[h_{JJ^{}}^{C;\mathrm{S}}]}\right)\hfill & |& \text{by (}\text{121}\text{)}\hfill \\ & =& \alpha _{J,\sigma (i)}\left(Q_L\right)\hfill & |& \text{by (}\text{122}\text{)}\hfill \\ & =& \alpha _{\sigma (i)}\hfill & |& \text{by (}\text{118}\text{),}\hfill \end{array}$$
i.e., $`\alpha ^{}=\sigma \alpha `$. In an analogous way, using (123), we obtain $`\xi ^{}=\xi `$.
To prove the converse implication, assume $`\xi ^{}=\xi `$ and let $`\sigma `$ be a permutation of $`1,\mathrm{},r`$ such that $`J^{}=\sigma J`$ and $`\alpha ^{}=\sigma \alpha `$. Due to Lemma 4.2, there exists $`D\mathrm{SU}n`$ such that $`\mathrm{SU}J^{}=D^1\mathrm{SU}JD`$. As shown in the proof of this lemma, $`D`$ can be chosen in such a way that (120) holds. Then (122) holds, too. We compute the characteristic classes of $`Q_LD`$:
$$\begin{array}{ccccc}\hfill \alpha _{J^{},i}\left(Q_LD\right)& =& \alpha _{J^{},i}\left(Q_L^{[h_{JJ^{}}^{D;\mathrm{S}}]}\right)\hfill & |& \text{by (}\text{119}\text{)}\hfill \\ & =& \alpha _{J,\sigma (i)}\left(Q_L\right)\hfill & |& \text{by (}\text{122}\text{)}\hfill \\ & =& \alpha _{\sigma (i)}\hfill & |& \text{by (}\text{118}\text{)}\hfill \\ & =& \alpha _i^{}\hfill & |& \text{by assumption}\hfill \\ & =& \alpha _{J^{},i}\left(Q_L^{}\right)\hfill & |& \text{by (}\text{118}\text{).}\hfill \end{array}$$
Analogously, using (123), we find $`\xi _J^{}\left(Q_LD\right)=\xi _J^{}\left(Q_L^{}\right)`$. It follows $`Q_LDQ_L^{}`$.
As suggested by Lemma 7.1, we introduce an equivalence relation on the set $`\mathrm{K}(P)`$: Let $`L,L^{}\mathrm{K}(P)`$, where $`L=(J;\alpha ,\xi )`$ and $`L^{}=(J^{};\alpha ^{},\xi ^{})`$. Write $`LL^{}`$ iff $`\xi ^{}=\xi `$ and there exists a permutation $`\sigma `$ of $`1,\mathrm{},r`$ such that $`J^{}=\sigma J`$ and $`\alpha ^{}=\sigma \alpha `$. The set of equivalence classes will be denoted by $`\widehat{\mathrm{K}}(P)`$.
###### Theorem 7.2
The assignment $`LQ_L`$ induces a bijection from $`\widehat{\mathrm{K}}(P)`$ onto $`\text{Howe}_{}(P)`$.
Proof: The assignment $`LQ_L`$ induces a map $`\mathrm{K}(P)\text{Howe}_{}(P)`$. This map is surjective by construction. Due to Lemma 7.1, it projects to $`\widehat{\mathrm{K}}(P)`$ and the projected map is injective.
With Theorem 7.2 we have accomplished the determination of $`\text{Howe}_{}(P)`$ and, therefore, of the set of orbit types $`\mathrm{OT}(๐^k,๐ข^{k+1})`$. Calculations for the latter can now be performed entirely on the level of the classifying set $`\widehat{\mathrm{K}}(P)`$.
## 8 Example: Gauge Orbit Types for $`\mathrm{SU2}`$
In this section, we are going to determine $`\mathrm{OT}(๐^k,๐ข^{k+1})`$ for an $`\mathrm{SU2}`$-gauge theory over the base manifolds discussed in Subsection 5.10. The set $`\mathrm{K}(2)`$ contains the elements
$$J_a=(1|2),J_b=(1,1|1,1),J_c=(2|1).$$
Here $`\mathrm{SU}J_a=\{\pm \mathrm{๐}\}_2`$ is the center, $`\mathrm{SU}J_b=\left\{\text{diag}(z,z^1)\right|z\mathrm{U1}\}\mathrm{U1}`$ is the toral subgroup and $`\mathrm{SU}J_c=\mathrm{SU2}`$. The strata corresponding to the elements of $`\mathrm{K}(P)_{J_a}`$, $`\mathrm{K}(P)_{J_b}`$, $`\mathrm{K}(P)_{J_c}`$ are, in the respective order, those with stabilizers isomorphic to $`\mathrm{SU2}`$, $`\mathrm{U1}`$, and the generic stratum. Accordingly, we shall refer to the first class as $`\mathrm{SU2}`$-strata and to the second class as $`\mathrm{U1}`$-strata. We have
$$\mathrm{K}(P)=\mathrm{K}(P)_{J_a}\mathrm{K}(P)_{J_b}\mathrm{K}(P)_{J_c}$$
(disjoint union). As we already know, $`\mathrm{K}(P)_{J_a}`$ is parametrized by $`\xi H^1(M,_2)`$ in case $`P`$ is trivial and is empty otherwise. Moreover, $`\mathrm{K}(P)_{J_c}`$ consists of $`P`$ itself. For $`\mathrm{K}(P)_{J_b}`$, variables are $`\alpha _i=1+\alpha _i^{(2)}`$, $`\alpha _i^{(2)}H^2(M,)`$, $`i=1,2`$. The system of equations (91), (92) reads
$`\alpha _1^{(2)}+\alpha _2^{(2)}`$ $`=`$ $`0`$ (124)
$`\alpha _1^{(2)}\alpha _2^{(2)}`$ $`=`$ $`c_2(P)`$ (125)
Using (124) to replace $`\alpha _2^{(2)}`$ in (125) we obtain
$$\alpha _1^{(2)}\alpha _1^{(2)}=c_2(P).$$
(126)
Note that here $`\alpha _1^{(2)}`$ is just the first Chern class of a reduction of $`P`$ to the subgroup $`\mathrm{U1}`$. According to this, Eq. (126) has been derived in the discussion of spontaneous symmetry breaking of $`\mathrm{SU2}`$ to $`\mathrm{U1}`$, see . Note also that when passing from $`\mathrm{K}(P)`$ to $`\widehat{\mathrm{K}}(P)`$, the pairs $`(\alpha _1,\alpha _2)`$ and $`(\alpha _2,\alpha _1)`$ label the same class of subbundles. Hence, solutions $`\alpha _1^{(2)}`$ of (126) have to be identified with their negative.
Let us now discuss Eq. (126) as well as the set $`\widehat{\mathrm{K}}(P)`$ in dependence of $`P`$ for some specific base manifolds $`M`$.
#### $`๐ด\mathbf{=}๐^\mathrm{๐}`$
Since $`H^2(M,)=0`$, Eq. (126) requires $`c_2(P)=0`$. Thus, in case $`P`$ is trivial, $`\widehat{\mathrm{K}}(P)`$ contains the $`_2`$-bundle, the $`\mathrm{U1}`$-bundle (both necessarily trivial) and $`P`$ itself. Accordingly, in the gauge orbit space there exist, besides the generic stratum, an $`\mathrm{SU2}`$-stratum and a $`\mathrm{U1}`$-stratum.
We remark that, due to the base manifold being a sphere, this result can be obtained much easier by homotopy arguments. It is, therefore, well known. The structure of the gauge orbit space in the present situation has been studied in great detail in . It has been shown that the two nongeneric strata of the gauge orbit space can be parametrized by means of an affine subspace of $`๐^k`$ which is acted upon by the Weyl group of $`\mathrm{SU2}`$ (the latter being just $`_2`$).
In case $`P`$ is nontrivial, both $`\mathrm{K}(P)_{J_b}`$ and $`\mathrm{K}(P)_{J_a}`$ are empty, so that $`\widehat{\mathrm{K}}(P)`$ contains only the bundle $`P`$ itself. Accordingly, there do not exist nongeneric strata in the gauge orbit space.
#### $`๐ด\mathbf{=}๐^\mathrm{๐}\mathbf{\times }๐^\mathrm{๐}`$
We write $`\alpha _1^{(2)}`$ in the form (105). Then Eq. (126) becomes
$$2ab\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}=c_2(P).$$
Thus, if $`P`$ is trivial then $`\widehat{\mathrm{K}}(P)`$ contains, in addition to $`P`$ itself, the $`_2`$-bundle, which is trivial, and the $`\mathrm{U1}`$-bundles labelled by $`a0`$, $`b=0`$ and $`a=0`$, $`b0`$, i.e., which are trivial over one of the $`2`$-spheres. The corresponding nongeneric strata in the gauge orbit space are one $`\mathrm{SU2}`$-stratum and infinitely many $`\mathrm{U1}`$-strata.
In case $`c_2(P)=2l\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}`$, $`l0`$, $`\widehat{\mathrm{K}}(P)`$ contains the $`\mathrm{U1}`$-bundles with $`a=q`$ and $`b=l/q`$, where $`q`$ is a (positive) divisor of $`m`$. Hence, here the nongeneric part of the gauge orbit space consists of finitely many $`\mathrm{U1}`$-strata.
Finally, in case $`c_2(P)=(2m+1)\gamma _{\mathrm{S}^2}^{(2)}\times \gamma _{\mathrm{S}^2}^{(2)}`$, $`\widehat{\mathrm{K}}(P)`$ contains only $`P`$ itself and the gauge orbit space consists only of the generic stratum.
#### $`๐ด\mathbf{=}๐^\mathrm{๐}`$
We use the notation introduced in Subsection 5.10. Writing $`\alpha _1^{(2)}`$ in the form (107) and using (106), Eq. (126) reads
$$2\left(a_{12}a_{34}a_{13}a_{24}+a_{14}a_{23}\right)\gamma _{\mathrm{T}^4}^{(4)}=c_2(P).$$
(127)
Hence, the result is similar to that for the case $`M=\mathrm{S}^2\times \mathrm{S}^2`$. The only difference is that, due to $`H^1(M,_2)_2^4`$, there exist $`16`$ different $`_2`$-bundles which are all contained in $`\widehat{\mathrm{K}}(P)`$ for $`P`$ being trivial. Accordingly, in this case the gauge orbit space contains $`16`$ $`\mathrm{SU2}`$-strata. Moreover, in case $`c_2(P)=2m\gamma _{\mathrm{T}^4}^{(4)}`$, $`m0`$, the number of solutions of (127) is infinite. Therefore, in this case there exist infinitely many $`\mathrm{U1}`$-strata in the gauge orbit space.
We remark that pure Yang-Mills theory on $`\mathrm{T}^4`$ has been discussed in , where the authors have studied the maximal Abelian gauge for gauge group $`\mathrm{SU}n`$. They have found that in case $`P`$ is nontrivial this gauge fixing is necessarily singular on Dirac strings joining magnetically charged defects in the base manifold. It is an interesting question whether there is a relation between such defects and the nongeneric strata of the gauge orbit space.
#### $`๐ด\mathbf{=}๐_๐^\mathrm{๐}\mathbf{\times }๐^\mathrm{๐}`$
Here (126) requires $`c_2(P)=0`$. Thus, in case $`P`$ is trivial $`\widehat{\mathrm{K}}(P)`$ contains, in addition to $`P`$ itself, the $`_2`$-bundles over $`\mathrm{L}_p^3\times \mathrm{S}^1`$, which are labelled by the elements of
$$H^1(\mathrm{L}_p^3\times \mathrm{S}^1,_2)\{\begin{array}{ccc}_2_2& |& p0\text{, even}\hfill \\ _2& |& \text{otherwise}\hfill \end{array}$$
as well as the $`\mathrm{U1}`$-bundles over $`\mathrm{L}_p^3\times \mathrm{S}^1`$, which are labelled by $`\alpha _1^{(2)}H^2(\mathrm{L}_p^3\times \mathrm{S}^1,)_p`$, modulo the identification $`\alpha _1^{(2)}\alpha _1^{(2)}`$. According to this, if $`p`$ is even, the nongeneric part of the gauge orbit space contains $`4`$ $`\mathrm{SU2}`$-strata and $`(p/2+1)`$ $`\mathrm{U1}`$-strata. If $`p`$ is odd, it contains $`2`$ $`\mathrm{SU2}`$-strata and $`((p+1)/2)`$ $`\mathrm{U1}`$-strata.
In case $`P`$ is nontrivial there do not exist nongeneric strata in the gauge orbit space.
## 9 Application: Kinematical Nodes in <br>$`2+1`$ Dimensional Chern-Simons Theory
Following , we consider Chern-Simons theory with gauge group $`\mathrm{SU}n`$ in the Hamiltonian approach. The Hamiltonian in Schrรถdinger representation is given by
$$H=\frac{\mathrm{\Lambda }}{2}_{\mathrm{\Sigma }_s}\frac{\mathrm{d}^2x}{\sqrt{h}}\mathrm{Tr}\left|\frac{\delta }{\delta A_\mu }+\frac{\mathrm{i}\mathrm{}}{4\pi }ฯต^{\mu \nu }A_\nu \right|^2+\frac{1}{4\mathrm{\Lambda }}_{\mathrm{\Sigma }_s}\mathrm{d}^2x\sqrt{h}\mathrm{Tr}\left(F_{\mu \nu }F^{\mu \nu }\right).$$
Geometrically, $`A_\mu `$ and $`F_{\mu \nu }`$ are the local representatives of a connection $`A`$ in a given $`\mathrm{SU}n`$-bundle $`P`$ over the $`2`$ dimensional space $`\mathrm{\Sigma }_s`$ and its curvature $`F`$, respectively. We assume $`\mathrm{\Sigma }_s`$ to be a Riemann surface of genus $`s`$. Note that, due to the dimension of $`\mathrm{\Sigma }_s`$, $`c_2(P)=0`$, i.e., $`P`$ is trivial. The configuration space is the gauge orbit space $`^k`$ associated to $`P`$. Physical states are given by cross sections of a complex line bundle $`\eta `$ over (the topological space) $`^k`$ with first Chern class $`c_1(\eta )=\mathrm{}`$. Correspondingly, they can be thought of as functionals $`\psi `$ on the space $`๐^k`$ of connections in $`P`$ which are subject to the Gauss law condition
$$_A^\mu \frac{\delta }{\delta A^\mu (x)}\psi (A)=\frac{\mathrm{i}\mathrm{}}{4\pi }ฯต^{\mu \nu }_\mu A_\nu (x)\psi (A).$$
(128)
Here $`_A`$ denotes the covariant derivative w.r.t. $`A`$. In it has been shown that if $`A`$ carries a nontrivial magnetic charge, i.e., if it can be reduced to some subbundle of $`P`$ with nontrivial first Chern class, all physical states obey $`\psi (A)=0`$. Such a connection is called a kinematical node. (Note that, due to $`\eta `$ being nontrivial, there exist also dynamical nodes, which differ from state to state.) The authors of argue that nodal gauge field configurations are relevant for the confinement mechanism. In the following we shall show that being a node is a property of strata. For that purpose, we reformulate the result of in our language.
###### Theorem 9.1
Let $`A๐^k`$ have orbit type $`[L]\widehat{\mathrm{K}}(P)`$, where $`L=(J;\alpha ,\xi )`$. If $`\alpha _i^{(2)}0`$ for some $`i`$ then $`A`$ is a kinematical node, i.e., $`\psi (A)=0`$ for all physical states $`\psi `$.
Proof: The proof follows the lines of . By assumption, $`A`$ can be reduced to a connection on an $`\mathrm{SU}J`$-subbundle $`QP`$ of class $`(\alpha _J(Q),\xi _J(Q))=(\alpha ,\xi )`$. $`\mathrm{\Sigma }_s`$ being a compact orientable $`2`$-manifold, $`H^2(\mathrm{\Sigma }_s,)`$. Hence, $`\alpha _i^{(2)}=c_i\gamma ^{(2)}`$, where $`c_i`$ and $`\gamma ^{(2)}`$ is a generator of $`H^2(\mathrm{\Sigma }_s,)`$. We define a $`1`$-parameter subgroup $`\{\stackrel{~}{\mathrm{\Phi }}_t|t\}`$ of $`\mathrm{SU}n`$ by
$$\stackrel{~}{\mathrm{\Phi }}_t=\left(\mathrm{exp}\left\{i\frac{c_1}{k_1}t\right\}\mathrm{๐}_{k_1}\mathrm{๐}_{m_1}\right)\mathrm{}\left(\mathrm{exp}\left\{i\frac{c_r}{k_r}t\right\}\mathrm{๐}_{k_r}\mathrm{๐}_{m_r}\right).$$
Using
$$\begin{array}{ccccc}\hfill (m_1c_1+\mathrm{}+m_rc_r)\gamma ^{(2)}& =& g\left(\stackrel{~}{m}_1\alpha _1^{(2)}+\mathrm{}+\stackrel{~}{m}_r\alpha _r^{(2)}\right)\hfill & & \\ & =& gE_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha )\hfill & |& \text{by (}\text{69}\text{)}\hfill \\ & =& g\beta _g(\xi )\hfill & |& \text{by (}\text{91}\text{)}\hfill \\ & =& 0\hfill & |& \text{by }g\beta _g=0\text{,}\hfill \end{array}$$
one checks $`det\stackrel{~}{\mathrm{\Phi }}_t=\mathrm{exp}\{\mathrm{i}t(m_1c_1+\mathrm{}+m_rc_r)\}=1`$. Commuting with $`\mathrm{SU}J`$ for all $`t`$, $`\stackrel{~}{\mathrm{\Phi }}_t`$ defines a $`1`$-parameter subgroup $`\{\mathrm{\Phi }_t|t\}`$ of $`๐ข^{k+1}`$ by
$$\mathrm{\Phi }_t(q)=\stackrel{~}{\mathrm{\Phi }}_tqQ,t.$$
Each element of this subgroup is constant on $`Q`$. Hence, so is the generator $`\varphi =\dot{\mathrm{\Phi }}(0)`$:
$$\varphi (q)=\mathrm{i}\left[\left(\frac{c_1}{k_1}\mathrm{๐}_{k_1}\mathrm{๐}_{m_1}\right)\mathrm{}\left(\frac{c_r}{k_r}\mathrm{๐}_{k_r}\mathrm{๐}_{m_r}\right)\right]qQ.$$
(129)
In particular,
$$_A\varphi =0.$$
(130)
According to this, for any state $`\psi `$,
$$_{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi _A^\mu \frac{\delta }{\delta A^\mu }\right)\psi (A)=_{\mathrm{\Sigma }_s}\mathrm{Tr}\left((_A^\mu \varphi )\frac{\delta }{\delta A^\mu }\right)\psi (A)=0.$$
For physical states, the Gauss law implies
$$_{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi \mathrm{d}A\right)\psi (A)=0.$$
(131)
Using (130), as well as the structure equation $`F=\mathrm{d}A+\frac{1}{2}[A,A]`$, we obtain
$$\begin{array}{ccc}\hfill _{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi \mathrm{d}A\right)& =& _{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi \mathrm{d}A_A\varphi A\right)\hfill \\ & =& _{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi \mathrm{d}A\mathrm{d}\varphi A[A,\varphi ]A\right)\hfill \\ & =& _{\mathrm{\Sigma }_s}\mathrm{Tr}\left(2\varphi \mathrm{d}A+\varphi [A,A]\right)\hfill \\ & =& 2_{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi F\right).\hfill \end{array}$$
(132)
Since $`A`$ is reducible to $`Q`$, $`F`$ has block structure $`\left(F_1\mathrm{๐}_{m_1}\right)\mathrm{}\left(F_r\mathrm{๐}_{m_r}\right)`$ with $`F_j`$ being $`(k_j\times k_j)`$-matrices. Thus, using (129),
$$_{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi F\right)=\mathrm{i}\underset{j=1}{\overset{r}{}}\frac{m_j}{k_j}c_j_{\mathrm{\Sigma }_s}\mathrm{Tr}F_j.$$
(133)
Now $`c_j`$ being just the first Chern classes of the elementary factors of the $`\mathrm{U}k_1\times \mathrm{}\times \mathrm{U}k_r`$-bundle $`Q^{[\mathrm{U}J]}`$, we have
$$_{\mathrm{\Sigma }_s}\mathrm{Tr}F_j=2\pi \mathrm{i}c_j,j=1,\mathrm{},r.$$
Inserting this into (133) and the latter into (132) we obtain
$$_{\mathrm{\Sigma }_s}\mathrm{Tr}\left(\varphi \mathrm{d}A\right)=4\pi \underset{j=1}{\overset{r}{}}\frac{m_j}{k_j}\left(c_j\right)^2.$$
Thus, in view of (131),
$$4\pi \underset{j=1}{\overset{r}{}}\frac{m_j}{k_j}\left(c_j\right)^2\psi (A)=0.$$
(134)
It follows that if one of the $`c_j`$ is nonzero then $`\psi (A)=0`$ for all physical states $`\psi `$, i.e., $`A`$ is a kinematical node. This proves the theorem.
Remark: Let us compare (134) with Formula $`(6)`$ in . Define $`k_i^{}=k_im_i`$ and $`m_i^{}=1`$. Then $`J^{}=(๐ค^{},๐ฆ^{})\mathrm{K}(n)`$ and $`\mathrm{U}J\mathrm{U}J^{}`$. Let $`\phi :\mathrm{U}J\mathrm{U}J^{}`$ denote the canonical embedding. Consider the subbundle $`Q^{[\phi ]}P`$. One can check that the elementary factors of this subbundle have first Chern classes $`c_i^{}=m_ic_i`$. Inserting $`k_i^{}`$, $`l_i^{}`$, and $`c_i^{}`$ into (134) one obtains Formula $`(6)`$ in . In fact, the authors of use that $`A`$ is reducible to $`Q^{}`$, rather than that it is even reducible to $`Q`$. This argument being โcoarserโ than ours, it still suffices to prove that any connection which is reducible to a subbundle with nontrivial magnetic charge is a kinematical node.
As a consequence of Theorem 9.1, one can speak of nodal and nonnodal strata. This information can be read off directly from the labels of the strata.
Let us discuss this in some more detail. Let $`J\mathrm{K}(n)`$ be given and consider Eqs. (91) and (92). Variables are $`\xi H^1(\mathrm{\Sigma }_s,_g)`$ and $`\alpha _i^{(2)}H^2(\mathrm{\Sigma }_s,)`$, $`i=1,\mathrm{},r`$. Since $`H^2(\mathrm{\Sigma }_s,)`$ is torsion-free, $`\beta _g`$ is trivial. Moreover, due to $`H^4(\mathrm{\Sigma }_s,)=0`$, $`a_ia_j`$ vanishes. Thus, the system of equations (91), (92) reduces to
$$E_{\stackrel{~}{๐ฆ}}^{(2)}(\alpha )=0.$$
(135)
Writing $`\alpha _i^{(2)}=c_i\gamma ^{(2)}`$ again, (135) becomes
$$\underset{i=1}{\overset{r}{}}\stackrel{~}{m}_ic_i=0.$$
The set of solutions of this equation is a subgroup $`G_{\stackrel{~}{๐ฆ}}^r`$. According to Theorem 9.1, the nonnodal strata are parametrized by $`\xi `$ and the neutral element of $`G_{\stackrel{~}{๐ฆ}}`$, whereas the nodal strata are labelled by $`\xi `$ and all the other elements of $`G_{\stackrel{~}{๐ฆ}}`$. For example, in the case of $`\mathrm{SU2}`$ we obtain the following.
$`J=(1|2)`$: Here $`G_{\stackrel{~}{๐ฆ}}=\{0\}`$, hence all strata are nonnodal.
$`J=(1,1|1,1)`$: We have $`G_{\stackrel{~}{๐ฆ}}=\{(c,c)|c\}^2`$. Since also $`\xi =0`$, each value of $`c`$ labels one stratum. That corresponding to $`c=0`$ is nonnodal, the others are nodal.
$`J=(2|1)`$: Here we have the generic stratum, which is nonnodal.
## 10 Summary
Starting from a principal $`\mathrm{SU}n`$-bundle $`P`$ over a compact connected orientable Riemannian $`4`$-manifold $`M`$, we have derived a classification of the orbit types of the action of the group of gauge transformations of $`P`$ on the space of connections in $`P`$. Orbit types are known to label the elements of the natural stratification, given by Kondracki and Rogulski , of the gauge orbit space associated to $`P`$. The interest in this stratification is due to the fact that the role of nongeneric strata in gauge physics is not clarified yet.
In order to accomplish the classification, we have utilized that orbit types are 1:1 with a certain class of subbundles of $`P`$ (called holonomy-induced Howe subbundles), factorized by the natural actions of vertical automorphisms of $`P`$ and of the structure group. We have shown that such classes of subbundles are labelled by symbols $`[(J;\alpha ,\xi )]`$, where
$$J=((k_1,\mathrm{},k_r),(m_1,\mathrm{},m_r))$$
is a pair of sequences of positive integers obeying
$$\underset{i=1}{\overset{r}{}}k_im_i=n,$$
$`\alpha =(\alpha _1,\mathrm{},\alpha _r)`$ where $`\alpha _iH^{}(M,)`$ are admissible values of the Chern class of $`\mathrm{U}k_i`$-bundles over $`M`$, and $`\xi H^1(M,_g)`$ with $`g`$ being the greatest common divisor of $`(m_1,\mathrm{},m_r)`$. The cohomology elements $`\alpha _i`$ and $`\xi `$ are subject to the relations
$`{\displaystyle \underset{i=1}{\overset{r}{}}}{\displaystyle \frac{m_i}{g}}\alpha _i^{(2)}`$ $`=`$ $`\beta _g(\xi ),`$
$`\alpha _1^{m_1}\mathrm{}\alpha _r^{m_r}`$ $`=`$ $`c(P),`$
where $`\beta _g:H^1(M,_g)H^2(M,)`$ is the connecting homomorphism associated to the short exact sequence of coefficient groups in cohomology
$$0_g0.$$
Finally, for any permutation $`\sigma `$ of $`\{1,\mathrm{},r\}`$, the symbols
$$\begin{array}{c}[((k_1,\mathrm{},k_r),(m_1,\mathrm{},m_r);(\alpha _1,\mathrm{},\alpha _r),\xi )],\\ [((k_{\sigma (1)},\mathrm{},k_{\sigma (r)}),(m_{\sigma (1)},\mathrm{},m_{\sigma (r)});(\alpha _{\sigma (1)},\mathrm{},\alpha _{\sigma (r)}),\xi )]\end{array}$$
have to be identified.
The result obtained enables one to determine which strata are present in the gauge orbit space, depending on the topology of the base manifold and the topological sector, i.e., the isomorphism class of $`P`$. For some examples we have discussed this dependence in detail. We have also shown that our result can be used to reformulate a sufficient condition on a connection to be a node for all physical states (when the latter are viewed as functionals of connections subject to the Gauss law). This condition has been derived in and applies to Chern-Simons Theory in $`2+1`$ dimensions.
In this way, our result may be viewed as one more step towards a systematic investigation of the physical effects related to nongeneric strata of the gauge orbit space.
We remark that orbit types still carry more information about the stratification structure. Namely, their partial ordering encodes how the strata are patched together in order to build up the gauge orbit space (cf. \[23, Thm. (4.3.5)\]). A derivation of this partial ordering will be published separately.
## Acknowledgements
The authors would like to thank C. Fleischhack, S. Kolb, A. Strohmaier, and S. Boller for interesting discussions. They are also grateful to T. Friedrich, J. Hilgert, and H.-B. Rademacher for useful suggestions, as well as to L.M. Woodward, C. Isham, and M. ฤadek who on request have been very helpful with providing specific information.
## Appendix A The Eilenberg-MacLane Spaces $`K(,2)`$ and $`K(_g,1)`$
In this appendix, we construct a model for each of the Eilenberg-MacLane spaces $`K(,2)`$ and $`K(_g,1)`$ and derive the integer-valued cohomology of these spaces. Consider the natural free action of $`\mathrm{U1}`$ on the sphere $`\mathrm{S}^{\mathrm{}}`$ which is induced from the natural action of $`\mathrm{U1}`$ on $`\mathrm{S}^{2n1}^n`$. The orbit space of this action is the complex projective space $`\mathrm{P}^{\mathrm{}}`$. Moreover, by viewing $`_g`$ as a subgroup of $`\mathrm{U1}`$, this action gives rise to a natural free action of $`_g`$ on $`\mathrm{S}^{\mathrm{}}`$. The orbit space of the latter is the lens space $`\mathrm{L}_g^{\mathrm{}}`$. By construction, one has principal bundles
$`\mathrm{U1}`$ $``$ $`\mathrm{S}^{\mathrm{}}\mathrm{P}^{\mathrm{}},`$ (136)
$`_g`$ $``$ $`\mathrm{S}^{\mathrm{}}\mathrm{L}_g^{\mathrm{}}.`$ (137)
Due to $`\pi _i(\mathrm{S}^{\mathrm{}})=0`$ $`i`$, the exact homotopy sequences induced by (136), (137) yield
$$\begin{array}{ccccc}\pi _i(\mathrm{P}^{\mathrm{}})& =& \pi _{i1}(\mathrm{U1})& =& \{\begin{array}{ccc}& |& i=2\hfill \\ 0& |& i2,\hfill \end{array}\hfill \\ \pi _i(\mathrm{L}_g^{\mathrm{}})& =& \pi _{i1}(_g)& =& \{\begin{array}{ccc}_g& |& i=1\hfill \\ 0& |& i=2,3,\mathrm{},\hfill \end{array}\hfill \end{array}$$
respectively. As a consequence, $`\mathrm{P}^{\mathrm{}}`$ is a model of $`K(,2)`$ and $`\mathrm{L}_g^{\mathrm{}}`$ is a model of $`K(_g,1)`$. In particular,
$$H^i(K(,2),)=H^i(\mathrm{P}^{\mathrm{}},)=\{\begin{array}{ccc}& |& i\text{ even}\hfill \\ 0& |& i\text{ odd},\hfill \end{array}$$
(138)
see \[8, Ch. VI, Prop. 10.2\], and
$$H^i(K(_g,1),)=H^i(\mathrm{L}_g^{\mathrm{}},)=\{\begin{array}{ccc}& |& i=0\hfill \\ _g& |& i0\text{, even}\hfill \\ 0& |& i0\text{, odd},\hfill \end{array}$$
(139)
see \[11, ยง24, p. 176\].
We notice that the vanishing of all homotopy groups of $`\mathrm{S}^{\mathrm{}}`$ also implies that the bundles (136) and (137) are universal for $`\mathrm{U1}`$ and $`_g`$, respectively. Hence, $`\mathrm{P}^{\mathrm{}}`$ and $`\mathrm{L}_g^{\mathrm{}}`$ are models of $`\mathrm{BU1}`$ and $`\mathrm{B}_g`$, respectively. For $`\mathrm{B}_g`$, this has been used in the proof of Lemma 5.9.
## Appendix B The Cohomology Algebra $`H^{}(\mathrm{BSU}J,)`$
Let $`J\mathrm{K}(n)`$, $`J=(๐ค,๐ฆ)`$. Recall from Corollary 5.8 that the cohomology algebra $`H^{}(\mathrm{BSU}J,)`$ is generated by the elements $`\gamma _{J,i}^{(2j)}`$, $`j=1,\mathrm{},k_i`$, $`i=1,\mathrm{},r`$, defined in (52). Here we are going to derive the relations these generators are subject to.
###### Proposition B.1
The generators $`\gamma _{J,i}^{(2j)}`$ of $`H^{}(\mathrm{BSU}J,)`$ are subject to the relation
$$E_๐ฆ^{(2)}\left(\gamma _J\right)=0.$$
(140)
Proof: Consider the homomorphism $`\left(\mathrm{B}j_J\right)^{}:H^{}(\mathrm{BU}J,)H^{}(\mathrm{BSU}J,)`$ induced by the embedding $`j_J:\mathrm{SU}J\mathrm{U}J`$. By construction, $`\gamma _{J,i}^{(2j)}=\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\gamma }_{J,i}^{(2j)}`$. Since there are no relations between the $`\stackrel{~}{\gamma }_{J,i}^{(2j)}`$, see Lemma 5.6, all the relations between the $`\gamma _{J,i}^{(2j)}`$ are derived from elements of the kernel of $`\left(\mathrm{B}j_J\right)^{}`$. We claim that
$$\mathrm{ker}\left(\mathrm{B}j_J\right)^{}=H^{}(\mathrm{BU}J,)E_๐ฆ^{(2)}\left(\stackrel{~}{\gamma }_J\right),$$
(141)
i.e., that $`\mathrm{ker}\left(\mathrm{B}j_J\right)^{}`$ is generated as an ideal by the single element $`E_๐ฆ^{(2)}\left(\stackrel{~}{\gamma }_J\right)`$. This implies the assertion.
In order to prove (141), consider the $`\mathrm{U1}`$-bundle $`\eta `$ given in (57). The Gysin sequence (58) induced by $`\eta `$ being exact, it yields
$$\mathrm{ker}\left(\mathrm{B}j_J\right)^{}=H^{}(\mathrm{BU}J,)c_1(\eta ).$$
Let us compute $`c_1(\eta )`$. We note that $`\eta `$ is induced as a principal bundle of the form (28) by the short exact sequence of Lie group homomorphisms
$$\mathrm{๐}\mathrm{SU}J\stackrel{j_J}{}\mathrm{U}J\stackrel{det_{\mathrm{U}n}i_J}{}\mathrm{U1}\mathrm{๐},$$
cf. (28). Hence, it has classifying map $`\mathrm{B}det_{\mathrm{U}n}\mathrm{B}i_J`$. Accordingly,
$$\begin{array}{ccccc}\hfill c_1(\eta )& =& \left(\mathrm{B}i_J\right)^{}\left(\mathrm{B}det_{\mathrm{U}n}\right)^{}\gamma _{\mathrm{U1}}^{(2)}\hfill & & \\ & =& \left(\mathrm{B}i_J\right)^{}\gamma _{\mathrm{U}n}^{(2)}\hfill & |& \text{by (}\text{78}\text{)}\hfill \\ & =& E_๐ฆ^{(2)}(\stackrel{~}{\gamma }_{J,1},\mathrm{},\stackrel{~}{\gamma }_{J,r})\hfill & |& \text{by (}\text{72}\text{)}\hfill \end{array}$$
This proves the proposition.
As a consequence of Proposition B.1, $`H^{}(\mathrm{BSU}J,)`$ is isomorphic to the polynomial ring
$$[x_{11},\mathrm{},x_{1k_1},\mathrm{},x_{r1},\mathrm{},x_{rk_r}]/(m_1x_{11}+\mathrm{}+m_rx_{r1}),$$
where $`\mathrm{deg}(x_{ij})=2j`$.
We remark that the relation (140) is a consequence of the relation between $`\delta _J`$ and $`\gamma _J`$ given in Theorem 5.13. Namely, using (71), Theorem 5.13, and $`g\beta _g=0`$ we find
$$E_๐ฆ^{(2)}\left(\gamma _J\right)=gE_{\stackrel{~}{๐ฆ}}^{(2)}\left(\gamma _J\right)=g\beta _g(\delta _J)=0.$$
This ensures, in particular, that (140) does not generate a relation independent of (84) on the level of the characteristic classes $`\alpha _J`$. (Nonetheless, we have already proved directly in Lemma 5.15 that there are no relations independent of (84).)
## Appendix C The Cohomology Algebra $`H^{}(\mathrm{BSU}J,_g)`$
Let $`J\mathrm{K}(n)`$, $`J=(๐ค,๐ฆ)`$. In this appendix, we derive $`H^{}(\mathrm{BSU}J,_g)`$. Recall that $`g`$ denotes the greatest common divisor of $`๐ฆ`$. Moreover, recall that $`\varrho _g:H^{}(\mathrm{BSU}J,)H^{}(\mathrm{BSU}J,_g)`$ denotes reduction modulo $`g`$. We start with some technical lemmas.
###### Lemma C.1
(a) Let $`\alpha H^{}(\mathrm{BSU}J,)`$ such that $`\varrho _g\left(\alpha \beta _g(\delta _J)\right)=0`$. Then $`\varrho _g(\alpha )=0`$ as well as $`\alpha \beta _g(\delta _J)=0`$.
(b) One has $`\mathrm{im}\beta _gH^{}(\mathrm{BSU}J,)\beta _g(\delta _J)`$.
(c) The homomorphism $`\varrho _g\beta _g:H^{2j+1}(\mathrm{BSU}J,_g)H^{2j+2}(\mathrm{BSU}J,_g)`$ is injective for $`j=0,1,2,\mathrm{}`$ .
Proof: (a) Let $`\alpha `$ be given as proposed. Then there exists $`\alpha ^{}H^{}(\mathrm{BSU}J,)`$ such that
$$\alpha \beta _g(\delta _J)=g\alpha ^{}.$$
(142)
Moreover, there exist $`\stackrel{~}{\alpha },\stackrel{~}{\alpha }^{}H^{}(\mathrm{BU}J,)`$ such that $`\alpha =\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }`$ and $`\alpha ^{}=\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }^{}`$. Due to Theorem 5.13 and (52),
$$\beta _g(\delta _J)=\left(\mathrm{B}j_J\right)^{}E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)$$
(143)
Using this, as well as (141), (142) implies that there exists $`\stackrel{~}{\alpha }^{\prime \prime }H^{}(\mathrm{BU}J,)`$ such that
$`\stackrel{~}{\alpha }E_{\stackrel{~}{๐ฆ}}\left(\stackrel{~}{\gamma }_J\right)`$ $`=`$ $`g\stackrel{~}{\alpha }^{}+\stackrel{~}{\alpha }^{\prime \prime }E_๐ฆ^{(2)}\left(\stackrel{~}{\gamma }_J\right)`$ (144)
$`=`$ $`g\left(\stackrel{~}{\alpha }^{}+\stackrel{~}{\alpha }^{\prime \prime }E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)\right).`$
Taking into account that $`H^{}(\mathrm{BU}J,)`$ is torsion-free, that the elements $`\stackrel{~}{\gamma }_{J,i}^{(2j)}`$ are generators, and that the greatest common divisor of the integers $`\stackrel{~}{m}_i`$ is $`1`$, from (144) we infer that there exists $`\stackrel{~}{\alpha }^{\prime \prime \prime }H^{}(\mathrm{BU}J,)`$ such that $`\stackrel{~}{\alpha }=g\stackrel{~}{\alpha }^{\prime \prime \prime }`$. Then $`\alpha =\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }=g\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }^{\prime \prime \prime }`$. It follows $`\varrho _g(\alpha )=0`$. Due to $`g\beta _g(\delta _J)=0`$, also $`\alpha \beta _g(\delta _J)=0`$, as asserted.
(b) Let $`\alpha \mathrm{im}\beta _g`$. By exactness of (153), $`g\alpha =0`$. Let $`\stackrel{~}{\alpha }H^{}(\mathrm{BU}J,)`$ such that $`\alpha =\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }`$. Then $`g\stackrel{~}{\alpha }\mathrm{ker}\left(\mathrm{B}j_J\right)^{}`$. According to (141), there exists $`\stackrel{~}{\alpha }^{}H^{}(\mathrm{BU}J,)`$ such that
$$g\stackrel{~}{\alpha }=\stackrel{~}{\alpha }^{}E_๐ฆ^{(2)}\left(\stackrel{~}{\gamma }_J\right)=g\left(\stackrel{~}{\alpha }^{}E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)\right).$$
Since $`H^{}(\mathrm{BU}J,)`$ is free Abelian, this implies $`\stackrel{~}{\alpha }=\stackrel{~}{\alpha }^{}E_{\stackrel{~}{๐ฆ}}^{(2)}\left(\stackrel{~}{\gamma }_J\right)`$. Using (143) we obtain $`\alpha =\left(\mathrm{B}j_J\right)^{}\stackrel{~}{\alpha }^{}\beta _g(\delta _J)`$. This shows (b).
(c) We note that (a) and (b) immediately imply that $`\mathrm{im}\beta _g\mathrm{ker}\varrho _g=\{0\}`$. Moreover, by exactness of (153), $`\beta _g`$ is injective in odd degree. This proves (c).
###### Lemma C.2
The homomorphism $`H^{2j}(\mathrm{BSU}J,_g)H^{2j+1}(\mathrm{BSU}J,_g)`$, $`\alpha \alpha \delta _J`$, is an isomorphism for $`j=0,1,2,\mathrm{}`$ .
Proof: First, we check injectivity. Let $`\alpha H^{2j}(\mathrm{BSU}J,_g)`$ such that $`\alpha \delta _J=0`$. Consider the homomorphism $`\varrho _g\beta _g`$. As an immediate consequence of the definition of $`\beta _g`$ (namely, as the connecting homomorphism in the long exact sequence (60)), this is a skew-derivation, i.e., for any $`\alpha _iH^{j_i}(\mathrm{BSU}J,_g)`$, $`i=1,2`$, one has
$$\varrho _g\beta _g(\alpha _1\alpha _2)=(\varrho _g\beta _g(\alpha _1))\alpha _2+(1)^{j_1}\alpha _1(\varrho _g\beta _g(\alpha _2)).$$
(145)
Hence, due to $`\beta _g`$ being trivial in even degree,
$$\varrho _g\beta _g(\alpha \delta _J)=\alpha \varrho _g\beta _g(\delta _J).$$
(146)
Since $`\varrho _g`$ is surjective in even degree, there exists $`\alpha ^{}H^{2j}(\mathrm{BSU}J,)`$ such that $`\alpha =\varrho _g\left(\alpha ^{}\right)`$. Inserting this on the rhs. of (146) we obtain
$$\varrho _g\beta _g(\alpha \delta _J)=\varrho _g\left(\alpha ^{}\beta _g(\delta _J)\right).$$
(147)
By assumption, the lhs. of (147) vanishes. Then Lemma C.1(a) implies $`\alpha =\varrho _g\left(\alpha ^{}\right)=0`$. This proves injectivity. To show surjectivity, let $`\alpha H^{2j+1}(\mathrm{BSU}J,_g)`$. Due to Lemma C.1(b), there exists $`\alpha ^{}H^{}(\mathrm{BSU}J,)`$ such that $`\beta _g(\alpha )=\alpha ^{}\beta _g(\delta _J)`$. Then
$$\begin{array}{ccc}\hfill \varrho _g\beta _g(\alpha )& =& \varrho _g\left(\alpha ^{}\beta _g(\delta _J)\right)\hfill \\ & =& \varrho _g\left(\alpha ^{}\right)\varrho _g\beta _g(\delta _J)\hfill \\ & =& \varrho _g\beta _g(\varrho _g\left(\alpha ^{}\right)\delta _J),\hfill \end{array}$$
where the last equality is due to (145) and the fact that $`\beta _g`$ is trivial in even degree. As a consequence, Lemma C.1(c) implies $`\alpha =\varrho _g\left(\alpha ^{}\right)\delta _J`$. This shows surjectivity and, therefore, concludes the proof of the lemma.
###### Lemma C.3
There holds $`\delta _J\delta _J=\{\begin{array}{ccc}0& |& g=2l+1\hfill \\ l\varrho _g\beta _g(\delta _J)& |& g=2l\hfill \end{array}`$.
Proof: We notice that both the collections of maps
$`\theta _1:H^1(,_g)`$ $``$ $`H^2(,_g),\alpha \alpha \alpha ,`$ (148)
$`\theta _2:H^1(,_g)`$ $``$ $`H^2(,_g),\alpha \varrho _g\beta _g(\alpha ),`$ (149)
define a natural transformation of cohomology functors for $`CW`$-complexes. Such transformations are called cohomology operations of type $`(1,_g;2,_g)`$ or, more generally, of type $`(i_1,\pi _1;i_2,\pi _2)`$ if they map $`H^{i_1}(,\pi _1)H^{i_2}(,\pi _2)`$ \[8, Ch. VII, Def. 12.2\]. Here $`\pi _1,\pi _2`$ are Abelian groups. One should note that cohomology operations need not consist of group homomorphisms. Nevertheless, the Abelian group structure of $`H^{i_2}(,\pi _2)`$ induces an according structure on the set of cohomology operations of type $`(i_1,\pi _1;i_2,\pi _2)`$. Due to a theorem of Serre \[8, Ch. VII, Thm. 12.3\], there exists a group isomorphism from the group so defined onto $`H^{i_2}(K(\pi _1,i_1),\pi _2)`$. This is given by evaluating the cohomology operations at some fixed characteristic element of $`H^{i_1}(K(\pi _1,i_1),\pi _1)`$.
To apply this theorem, we choose $`\mathrm{L}_g^{\mathrm{}}`$ as a model of $`K(_g,1)`$ and $`\delta _g`$ as a characteristic element of $`H^1(K(_g,1),_g)`$. Assume, for a moment, that there holds
$$\theta _1(\delta _g)=\{\begin{array}{ccc}0& |& g=2l+1\hfill \\ l\theta _2(\delta _g)& |& g=2l.\hfill \end{array}$$
(150)
Then the cohomology operations are related by
$$\theta _1=\{\begin{array}{ccc}0& |& g=2l+1\hfill \\ l\theta _2& |& g=2l.\hfill \end{array}$$
Thus, in order to prove the lemma, we have to show (150). Although the cohomology of the spaces $`\mathrm{L}_g^{\mathrm{}}`$ is well known, this particular relation can rarely be found in textbooks. An exception is the case $`g=2`$, where $`\mathrm{L}_2^{\mathrm{}}=\mathrm{P}^{\mathrm{}}`$. For the sake of completeness, in the following we derive (150) for general $`g`$ from the case $`g=2`$.
One can show that $`H^2(\mathrm{L}_g^{\mathrm{}},_g)_g`$ and that it is generated by $`\varrho _g\beta _g(\delta _g)`$, where $`\delta _g`$ is a generator of $`H^1(\mathrm{L}_g^{\mathrm{}},_g)`$ . Thus, there exists $`a_g`$ such that
$$\theta _1(\delta _g)=\delta _g\delta _g=a\varrho _g\beta _g(\delta _g).$$
Since $`2\delta _g\delta _g=0`$, $`2a=0`$. Thus, if $`g`$ is odd then $`a=0`$. If $`g`$ is even then either $`a=l`$, where $`g=2l`$, or $`a=0`$. To rule out the second case, it suffices to find $`\delta H^1(\mathrm{L}_g^{\mathrm{}},_g)`$ such that $`\delta \delta 0`$. Consider the composite homomorphism
$$H^1(\mathrm{L}_g^{\mathrm{}},_g)\stackrel{\varrho _2}{}H^1(\mathrm{L}_g^{\mathrm{}},_2)\stackrel{p^{}}{}H^1(\mathrm{P}^{\mathrm{}},_2).$$
(151)
Here $`\varrho _2`$ is reduction modulo $`2`$ and $`p`$ is the projection in the principal bundle
$$_l\mathrm{P}^{\mathrm{}}\stackrel{p}{}\mathrm{L}_g^{\mathrm{}}$$
which arises by factorizing $`\mathrm{P}^{\mathrm{}}=\mathrm{S}^{\mathrm{}}/_2`$ by the residual action of $`_g/_2_l`$. We check that (151) is surjective. For the part of $`\varrho _2`$ this is obvious, due to $`g`$ being even. For the part of $`p^{}`$, we note that, by virtue of the Hurewicz and the Universal Coefficient Theorem, $`p^{}`$ is the $`_2`$-dual of the homomorphism $`p_{}:\pi _1\mathrm{P}^{\mathrm{}}\mathrm{L}_g^{\mathrm{}}`$. The latter is easily seen to be injective. Therefore, $`p^{}`$ is surjective, as asserted. As a consequence, there exists $`\delta H^1(\mathrm{L}_g^{\mathrm{}},_g)`$ such that $`p^{}\varrho _2(\delta )=\delta _2`$, where $`\delta _2`$ denotes the generator of $`H^1(\mathrm{P}^{\mathrm{}},_2)`$. We compute
$$p^{}\varrho _2(\delta \delta )=(p^{}\varrho _2(\delta ))(p^{}\varrho _2(\delta ))=\delta _2\delta _2.$$
Here the rhs. is known to be nontrivial \[8, Ch. VI, Prop. 10.2\]. Hence, $`\delta \delta `$ must have been nontrivial. This proves the Lemma.
###### Proposition C.4
$`H^{}(\mathrm{BSU}J,_g)`$ is generated over $`_g`$ by the elements $`\delta _J`$ and $`\varrho _g\left(\gamma _{J,i}^{(2j)}\right)`$, $`j=1,\mathrm{},k_i`$, $`i=1,\mathrm{},r`$. The generators are subject to the relation
$$\delta _J\delta _J=\{\begin{array}{ccc}0& |& g=2l+1\hfill \\ lE_{\stackrel{~}{๐ฆ}}^{(2)}\left(\gamma _J\right)& |& g=2l\hfill \end{array}$$
(152)
Proof: Consider the long exact sequence (60) for the space $`\mathrm{BSU}J`$. Since the integer-valued cohomology of $`\mathrm{BSU}J`$ is trivial in odd degree, this sequence splits into exact portions
$$\begin{array}{c}0\stackrel{\varrho _g}{}H^{2j+1}(\mathrm{BSU}J,_g)\stackrel{\beta _g}{}H^{2j+2}(\mathrm{BSU}J,)\hfill \\ \stackrel{\mu _g}{}H^{2j+2}(\mathrm{BSU}J,)\stackrel{\varrho _g}{}H^{2j+2}(\mathrm{BSU}J,_g)\stackrel{\beta _g}{}0,\hfill \end{array}$$
(153)
where $`j=0,1,2,\mathrm{}`$ . As an immediate consequence, $`\varrho _g`$ is surjective in even degree. Thus, $`H^{\mathrm{even}}(\mathrm{BSU}J,_g)`$ is generated by $`\varrho _g\left(\gamma _{J,i}^{(2j)}\right)`$, $`j=1,\mathrm{},k_i`$, $`i=1,\mathrm{},r`$. Then Lemma C.2 implies that the whole of $`H^{}(\mathrm{BSU}J,_g)`$ is generated by these elements together with $`\delta _J`$.
It remains to determine the relations the generators are subject to. First, consider relations among the generators $`\varrho _g\left(\gamma _{J,i}^{(2j)}\right)`$. These arise from the relation (140) and from the elements of $`\mathrm{ker}\varrho _g`$. Due to $`E_๐ฆ^{(2)}=gE_{\stackrel{~}{๐ฆ}}^{(2)}`$, the $`\text{mod}g`$-reduction of (140) is trivially satisfied. Moreover, relations generated by elements of $`\mathrm{ker}\varrho _g`$ are already taken into account by taking $`_g`$ as the base ring. Thus, there is no relation among the generators $`\varrho _g(\gamma _J,i^{(2j)})`$.
Next, consider relations involving $`\delta _J`$. One such relation is provided by Lemma C.3. It covers all relations in even degree, because the latter must contain an even power of $`\delta _J`$, hence can be written without $`\delta _J`$. Relations in odd degree, on the other hand, are of the form $`\alpha \delta _J=0`$, where $`\alpha `$ is of even degree. Due to Lemma C.2, then $`\alpha =0`$. Consequently, there are no further relations.
Due to Proposition C.4, if $`g`$ is odd then $`H^{}(\mathrm{BSU}J,_g)`$ is isomorphic to the polynomial ring
$$_g[x,x_{11},\mathrm{},x_{1k_1},\mathrm{},x_{r1},\mathrm{},x_{rk_r}],$$
whereas if $`g`$ is even then it is isomorphic to
$$_g[x,x_{11},\mathrm{},x_{1k_1},\mathrm{},x_{r1},\mathrm{},x_{rk_r}]/\left(x^2\left(\stackrel{~}{m}_1x_{11}+\mathrm{}+\stackrel{~}{m}_rx_{r1}\right)\right),$$
where $`g=2l`$. Here $`\mathrm{deg}(x)=1`$ and $`\mathrm{deg}(x_{ij})=2j`$. |
warning/0003/astro-ph0003365.html | ar5iv | text | # Cold and Fuzzy Dark Matter
\[
## Abstract
Cold dark matter (CDM) models predict small-scale structure in excess of observations of the cores and abundance of dwarf galaxies. These problems might be solved, and the virtues of CDM models retained, even without postulating ad hoc dark matter particle or field interactions, if the dark matter is composed of ultra-light scalar particles ($`m10^{22}`$eV), initially in a (cold) Bose-Einstein condensate, similar to axion dark matter models. The wave properties of the dark matter stabilize gravitational collapse providing halo cores and sharply suppressing small-scale linear power.
\]
Introduction.โ Recently, the small-scale shortcomings of the otherwise widely successful cold dark matter (CDM) models for structure formation have received much attention (see and references therein). CDM models predict cuspy dark matter halo profiles and an abundance of low mass halos not seen in the rotation curves and local population of dwarf galaxies respectively. Though the significance of the discrepancies is still disputed and solutions involving astrophysical processes in the baryonic gas may still be possible (e.g. ), recent attention has mostly focused on solutions involving the dark matter sector.
In the simplest modification, warm dark matter ($`m\mathrm{keV}`$) replaces CDM and suppresses small-scale structure by free-streaming out of potential wells , but this modification may adversely affect structure at somewhat larger scales. Small-scale power could be suppressed more cleanly in the initial fluctuations, perhaps originating from a kink in the inflaton potential , but its regeneration through non-linear gravitational collapse would likely still produce halo cusps .
More radical suggestions include strong self-interactions either between dark matter particles or in the potential of axion-like scalar field dark matter . While interesting, these solutions require self-interactions wildly in excess of those expected for weakly interacting massive particles or axions respectively.
In this Letter, we propose a solution involving free particles only. The catch is that the particles must be extraordinarily light ($`m10^{22}`$eV) so that their wave nature is manifest on astrophysical scales. Under this proposal, dark matter halos are stable on small scales for the same reason that the hydrogen atom is stable: the uncertainty principle in wave mechanics. We call this dark matter candidate fuzzy cold dark matter (FCDM).
Equations of Motion.โ It is well known that if the dark matter is composed of ultra-light scalar particles $`m1`$eV, the occupation numbers in galactic halos are so high that the dark matter behaves as a classical field obeying the wave equation
$$\mathrm{}\varphi =m^2\varphi ,$$
(1)
where we have set $`\mathrm{}=c=1`$. On scales much larger than the Compton wavelength $`m^1`$ but much smaller than the particle horizon, one can employ a Newtonian approximation to the gravitational interaction embedded in the covariant derivatives of the field equation and a non-relativistic approximation to the dispersion relation. It is then convenient to define the wavefunction $`\psi Ae^{i\alpha }`$, out of the amplitude and phase of the field $`\varphi =A\mathrm{cos}(mt\alpha )`$, which obeys
$$i(_t+\frac{3}{2}\frac{\dot{a}}{a})\psi =(\frac{1}{2m}^2+m\mathrm{\Psi })\psi ,$$
(2)
where $`\mathrm{\Psi }`$ is the Newtonian gravitational potential. For the unperturbed background, the right hand side vanishes and the energy density in the field, $`\rho =m^2|\psi |^2/2`$, redshifts like matter $`\rho a^3`$.
On time scales short compared with the expansion time, the evolution equations become
$`i_t\psi =({\displaystyle \frac{1}{2m}}^2+m\mathrm{\Psi })\psi ,^2\mathrm{\Psi }=4\pi G\delta \rho .`$ (3)
Assuming the dark matter also dominates the energy density, we have $`\delta \rho =m^2\delta |\psi |^2/2`$. This is simply the non-linear Schrรถdinger equation for a self-gravitating particle in a potential well. In the particle description, $`\psi `$ is proportional to the wavefunction of each particle in the condensate.
Jeans / de Broglie Scale.โ The usual Jeans analysis tells us that when gravity dominates there exists a growing mode $`e^{\gamma t}`$ where $`\gamma ^2=4\pi G\rho `$; however a free field oscillates as $`e^{iEt}`$ or $`\gamma ^2=(k^2/2m)^2`$. In fact, $`\gamma ^2=4\pi G\rho (k^2/2m)^2`$ and therefore there is a Jeans scale
$`r_J`$ $`=`$ $`2\pi /k_J=\pi ^{3/4}(G\rho )^{1/4}m^{1/2},`$ (4)
$`=`$ $`55m_{22}^{1/2}(\rho /\rho _b)^{1/4}(\mathrm{\Omega }_mh^2)^{1/4}\mathrm{kpc},`$ (5)
below which perturbations are stable and above which they behave as ordinary CDM. Here $`m_{22}=m/10^{22}`$eV and $`\rho _b=2.8\times 10^{11}\mathrm{\Omega }_mh^2M_{}`$ Mpc<sup>-3</sup> is the background density. The Jeans scale is the geometric mean between the dynamical scale and the Compton scale (c.f. ) as originally shown in a more convoluted manner by .
The existence of the Jeans scale has a natural interpretation: it is the de Broglie wavelength of the ground state of a particle in the potential well. To see this, note that the velocity scales as $`v(G\rho )^{1/2}r`$ so that the de Broglie wavelength $`\lambda (mv)^1m^1(G\rho )^{1/2}r^1`$. Setting $`r_J=\lambda =r`$, returns the Jeans scale. Stability below the Jeans scale is thus guaranteed by the uncertainty principle: an increase in momentum opposes any attempt to confine the particle further.
The physical scale depends weakly on the density, but in a dark matter halo $`\rho `$ will be much larger than the background density $`\rho _b`$. Consider the density profile of a halo of mass $`M`$ $`[(4\pi r_v^3/3)200\rho _b`$, in terms of the virial radius $`r_v`$\] found in CDM simulations
$`\rho (r,M)`$ $``$ $`{\displaystyle \frac{200}{3}}{\displaystyle \frac{f\rho _b}{(cr/r_v)(1+cr/r_v)^2}},`$ (6)
where $`f(c)=c^3/[\mathrm{ln}(1+c)c/(1+c)]`$ and the concentration parameter $`c`$ depends weakly on mass. This profile implies an $`r^1`$ cusp for $`r<r_v/c`$ which will be altered by the presence of the Jeans scale. Solving for the Jeans scale in the halo $`r_{J\mathrm{h}}`$ as a function of its mass using the enclosed mean density yields
$$r_{J\mathrm{h}}3.4(c_{10}/f_{10})^{1/3}m_{22}^{2/3}M_{10}^{1/9}(\mathrm{\Omega }_mh^2)^{2/9}\mathrm{kpc},$$
(7)
where we have scaled the mass dependent factors to the regime of interest $`c_{10}=c/10`$, $`f_{10}=f(c)/f(10)`$, and $`M_{10}=M/10^{10}M_{}`$. For estimation purposes, we have assumed $`r_{J\mathrm{h}}r_v/c`$ which is technically violated for $`M_{10}1`$ and $`m=10^{22}`$ eV, but with only a mild effect. In the smallest halos, the Jeans scale is above the turnover radius $`r_v/c`$, and there is no region where the density scales as $`r^1`$. The maximum circular velocity will then be lower than that implied by eqn. (6). More massive halos will have their cuspy $`r^1`$ behavior extend from $`r=r_v/c`$ down to $`r_{J\mathrm{h}}`$.
These simple scalings show that the wave nature of dark matter can prevent the formation of the kpc scale cusps and substructure in dark matter halos if $`m10^{22}`$eV. However, alone they do not determine what does form instead. To answer this question, cosmological simulations will be required and this lies beyond the scope of the present work. Instead we provide here the tools necessary to perform such a study, a discussion of possible astrophysical implications and illustrative one dimensional simulations comparing FCDM and CDM.
Linear Perturbations.โ The evolution of fluctuations in the linear regime provides the initial conditions for cosmological simulations and also directly affects the abundance of dark matter halos. Because the initial conditions are set while the fluctuations are outside the horizon, we must generalize the Newtonian treatment above to include relativistic effects.
Following the โgeneralized dark matterโ (GDM) approach of , we remap the equations of motion for the the scalar field in equation (1) onto the continuity and Euler equations of a relativistic imperfect fluid. First, note that in the Newtonian approximation, the current density $`๐ฃ\psi ^{}\psi \psi \psi ^{}`$ plays the role of momentum density so that โprobability conservationโ becomes the continuity equation. The dynamical aspect of equation (2) then becomes the Euler equation for a fluid with an effective sound speed $`c_{\mathrm{eff}}^2=k^2/4a^2m^2`$, where $`k`$ is the comoving wavenumber.
This Newtonian relation breaks down below the Compton scale which for any mode will occur when $`a<k/2m`$. In this regime, the scalar field is slowly-rolling in its potential rather than oscillating and it behaves like a fluid with an effective sound speed $`c_{\mathrm{eff}}^2=1`$ . For our purposes, it suffices to simply join these asymptotic solutions and treat the FCDM as GDM with
$$c_{\mathrm{eff}}^2=\{\begin{array}{cc}1,\hfill & ak/2m\text{ ,}\hfill \\ k^2/4a^2m^2,\hfill & a>k/2m\text{ ,}\hfill \end{array}$$
(8)
with no anisotropic stresses in linear theory. We have verified that the details of this matching have a negligible effect on the results. Since the underlying treatment is relativistic, this prescription yields a consistent, covariant treatment of the dark matter inside and outside the horizon. The linear theory equations including radiation and baryons are then solved in the usual way but with initial curvature perturbations in the radiation and no perturbations in the FCDM .
The qualitative features of the solutions are easily understood. The comoving Jeans wavenumber scales with the expansion as $`k_Ja\rho _b^{1/4}(a)`$ or $`a^{1/4}`$ during matter domination and constant during radiation domination. Because the comoving Jeans scale is nearly constant, perturbation growth above this scale generates a sharp break in the spectrum. More precisely, the critical scale is $`k_J`$ at matter-radiation equality $`k_{J\mathrm{eq}}=9m_{22}^{1/2}\mathrm{Mpc}^1.`$ Numerically, we find that the linear density power spectrum of FCDM is suppressed relative to the CDM case by
$$P_{\mathrm{FCDM}}(k)=T_\mathrm{F}^2(k)P_{\mathrm{CDM}}(k),T_\mathrm{F}(k)\frac{\mathrm{cos}x^3}{1+x^8},$$
(9)
where $`x=1.61m_{22}^{1/18}k/k_{J\mathrm{eq}}`$. The power drops by a factor of 2 at
$$k_{1/2}\frac{1}{2}k_{J\mathrm{eq}}m_{22}^{1/18}=4.5m_{22}^{4/9}\mathrm{Mpc}^1.$$
(10)
The break in $`k`$ is much sharper than those expected from inflation or quartic self-interaction of a scalar field. In the latter, the Jeans scale is fixed in physical coordinates so that the suppression is spread over 3-4 orders of magnitude in scale .
Low Mass Halos.โ In the CDM model, the abundance of low mass halos is too high when compared with the luminosity function of dwarf galaxies in the Local Group . Based on analytic scalings, Kamionkowski & Liddle argued that a sharp cutoff in the initial power spectrum at $`k=4.5h\mathrm{Mpc}^1`$ might solve this problem. Thus, the FCDM cutoff at $`k4.5\mathrm{Mpc}^1`$, produced if $`m10^{22}`$eV is chosen to remove kpc scale cusps, may solve the low mass halo abundance problem as well. Whether the required masses actually coincide in detail can only be addressed by simulations.
Numerical simulations of CDM with a smooth cutoff in the initial power spectrum qualitatively confirm the analytic estimates but suggest that a somewhat larger scale may be necessary: half power at $`k=2h\mathrm{Mpc}^1`$ reduces the $`z=3`$ abundance of $`10^{10}h^1M_{}`$ halos by a factor of $`5`$ and the abundance of $`10^{11}h^1M_{}`$ halos by a factor of $`3`$ . Note, however, that our model produces a much sharper cutoff in the power spectrum than in the model tested. Furthermore, astrophysical influences such as feedback or photoionization may have prevented dwarf galaxy halos from accumulating much gas or stars .
First Objects and Reionization.โ At very high redshift, much of the star formation in a CDM model is predicted to occur in low-mass halos which are not present in the FCDM model. In a CDM model the first round of star formation is thought to occur in objects of mass $`10^5M_{}`$ (see and references therein) due to molecular hydrogen cooling. The consequent destruction of molecular hydrogen implies that it is larger mass objects $`10^8M_{}`$, where atomic cooling is possible, that are responsible for reionization. In our scenario, if the cutoff scale in eqn. (10) were set to reduce the abundance of $`M10^9M_{}`$ halos, reionization could be delayed and the number of detectable galaxies prior to reionization reduced by a factor of $`5`$ .
Live Halos.โ Precisely what effect the Jeans (de Broglie) scale has on the structure and abundance of low mass halos is best answered through simulations. To provide some insight on these issues we conclude with simulations of the effects in one dimension.
We solve the wave equation (3) in an interval $`0<x<L`$ with boundary $`\psi (0)=\psi (L)=0`$. At $`t=0`$, the density perturbation is $`\delta \rho =\rho _0\mathrm{sin}(\pi x/L)\rho _b`$, with $`\psi `$ real. We define the Jeans length $`r_J`$ by equation (4) with the density $`\rho _0`$; then the choice of $`r_J/L`$ specifies $`m`$. It is also convenient to define the dynamical time scale $`t_{\mathrm{dyn}}=(4\pi G\rho _0)^{1/2}`$. A one dimensional CDM simulation with the same initial density and zero initial velocity was run for comparison.
For $`r_JL`$, the field model does not form a gravitating halo. For $`r_JL`$ a gravitationally bound halo is formed, but the cusp, which is clearly seen in the CDM simulation, is not observed (see Fig. 1). Interference effects cause continuous evolution on the dynamical time scale $`t_{\mathrm{dyn}}`$ (c.f. ). Note however that the gravitational acceleration is much smoother so that the trajectories of test particles (i.e. visible matter) will be less affected by these wiggles.
For $`r_JL`$, the density computed from the field follows the CDM simulation when smoothed over many Jeans scales. This is to be expected because in this limit the Schrรถdinger equation can be solved in the geometrical optics approximation. Nonetheless, interference features localized in space $`(\delta xr_J)`$ and time $`(\delta tt_{\mathrm{dyn}})`$ are quite strong, of order 1. Again these small-scale features make only a small contribution to the gravitational acceleration.
Discussion.โ We have shown that the wave properties of ultra light ($`m10^{22}`$eV) dark matter can suppress kpc scale cusps in dark matter halos and reduce the abundance of low mass halos. While such a mass may seem unnatural from a particle physics standpoint , even lighter scalar fields ($`m10^{33}`$ eV, where the wave nature is manifest across the whole particle horizon) have been proposed to provide a smooth energy component to explain observations of accelerated expansion .
Three dimensional numerical simulations are required to determine whether our proposal works in detail. Our one dimensional simulations suggest that the small-scale cutoff appears at $`rr_J`$ and that the density profile on these scales not only is not universal but also evolves continuously on the dynamical time scale (or faster) due to interference effects. The observable rotation curves are smoother than the density profile so that this prediction, while testable with high-resolution data, is not obviously in conflict with the data today. Likewise, the time-variation of the potential is smaller than that of the density but can in principle transfer energy from the FCDM to the baryons in the halo. This could puff up the baryons in dwarf galaxies while bringing the FCDM closer to a stationary ground state, but the precise evolution requires detailed calculations.
Our Jeans scale is a weak function of density, $`r_J\rho ^{1/4}`$. This has two testable consequences. The first is a sharp cutoff at $`k4.5m_{22}^{1/2}\mathrm{Mpc}^1`$ in the linear power. Quantities related to the abundance of low mass halos, e.g., dwarf galaxies in the Local Group, the first objects, faint galaxies at very high redshifts, and reionization can be seriously affected by the cutoff. Counterintuitively, quantities related to the non-linear power spectrum of the dark matter are only weakly affected due to the gravitational regeneration of small-scale power . The second consequence is that choosing the mass to set the core radius in one class of dark matter dominated objects sets the core radius in another set, given the ratio of characteristic densities. This relation can in principle be tested by comparing the local dwarf spheroidals to higher mass systems.
While the detailed implications remain to be worked out, fuzzy cold dark matter provides an interesting means of suppressing the excess small-scale power that plagues the cold dark matter scenario.
Acknowledgements.โ We thank J.P. Ostriker, P.J.E. Peebles, D.N. Spergel, M. Srednicki, and M. White for useful discussions. WH and AG are supported by the Keck Foundation and NSF-9513835; RB by Institute Funds and the NSF under Grant No. PHY94-07194. |
warning/0003/hep-th0003076.html | ar5iv | text | # Holographic RG and Cosmology in Theories with Quasi-Localized Gravity
## 1 Introduction
Randall and Sundrum (RS) have recently shown that it is possible to localize gravity to a brane in five dimensional anti-de Sitter space . In this model the theory on the brane reproduces four dimensional Einstein gravity at large distances even though the size of the extra dimension is infinitely large. This idea has sparked a flurry of activity in this field (see for related earlier work). Gregory, Rubakov and Sibiryakov (GRS) proposed a modified version of the RS model in which gravity appears five dimensional both at short and at long distances (see for a related idea), while at intermediate scales four dimensional gravity is reproduced. The GRS model has a positive tension brane, as in the RS model, and negative tension branes at a large proper distance away from the positive tension brane on either side of it. These negative tension branes have half the tension of the central brane. The brane tensions and the negative bulk cosmological constant are tuned such that the background is static, and the geometry between the branes is a slice of AdS<sub>5</sub>, while beyond the negative tension branes the spacetime is ordinary 5D Minkowski space. The fact that 4D gravity is reproduced at intermediate scales has been explained in , and the reason for this is that in this model the 4D graviton is replaced by a resonance with a finite lifetime, which can decay into the bulk; thus gravity is only quasi-localized. It was also suggested in that there may be a connection in these theories to bulk supersymmetry and vanishing of the cosmological constant. It has been shown in , that at intermediate distance scales the theory indeed reproduces the results of ordinary general relativity due to the bending of the brane in the presence of inhomogenous matter on the brane. The reason for this is that the effect of the bending of the brane exactly cancels the effects of the extra polarization in the massive graviton propagator (up to corrections that can be made arbitrary small by adjusting the width of the resonance).
We should stress that one of the essential features of these models is that they do not have a 4D low-energy effective field theory description (see also ). Instead of an effective 4D theory, these models have a an effective 5D theory at large distances which can be derived through a holographic renormalization group (RG) flow. This holographically renormalized theory will play the role of the low-energy effective theory. In order to simply perform a calculation at a given energy scale on the brane one performs the RG running to that scale in the theory on the brane. From the conjectured AdS/CFT correspondence the RG flow corresponds to moving the brane a finite distance into the bulk. This procedure is referred to as the holographic RG and will be the key to understanding the theory at large distances. At intermediate energies, the RG flow corresponds to moving the brane inside the AdS slice. Since an exactly AdS bulk corresponds to a conformal field theory, the brane tension of the effective theory remain unchanged. However, at distances large enough that effectively the branes have crossed, the effective low-energy theory will be that of a tensionless brane in 5D Minkowski space. This procedure can also be implemented for models which are smooth versions of the GRS model (see ), with the difference that there will be a continuous running in the brane tension of the theory. However, as long as the asymptotic metrics are equivalent to those of GRS the asymptotic form of the low energy effective theory will be the same as for GRS. In fact, a detailed analysis below will show that one finds that at intermediate distances (more precisely at a scale $`k^1e^{ky_0}`$, where $`y_0`$ is the location of the negative tension brane in the GRS model) the model becomes equivalent to the model recently proposed by Dvali, Gabadaze and Porrati , where a tensionless brane is embedded into 5D Minkowski space, but there is an additional induced four dimensional curvature term on the brane present. The induced operator is a consequence of the holographic renormalization. In addition, the radion mode of the GRS model (which corresponds to fluctuations of the distance between the two branes) will also be localized on the tensionless brane, as predicted in . This radion field will have a wrong-signed kinetic term, which is needed to cancel the effects of the extra graviton polarization in the DGP model. However, at very long distances, where the graviton mode becomes 5 dimensional, the radion will start to dominate and give rise to a peculiar 4D scalar antigravity, as discussed in . Due to the negative kinetic term of the radion these theories are probably not internally consistent at large scales; but from a purely phenomenological point of view it is still interesting to use these results to see how the cosmology of these models deviates from the ordinary FRW expansion of the Universe at large scales.
The paper is organized as follows: in Section 2 we explain the basic idea behind the holographic renormalization group, and calculate the effective brane tension and induced curvature term on the brane. In Section 3 we review the calculation of the induced radion kinetic term on the effective brane. In Section 4 we use the effective holographic theory obtained in Sections 2 and 3 to calculate the graviton propagator at large distances. We speculate on the cosmology of these models in Section 5, and conclude in Section 6.
## 2 Holographic Renormalization in Quasi-Localized Gravity Scenarios
Consider a 5D metric of the form
$$ds^2=dy^2+e^{A(y)}\eta _{\mu \nu }dx^\mu dx^\nu ,$$
(1)
where the warp factor $`A(y)`$ approaches the AdS form for $`|y|y_0`$:
$$A(y)2k|y|,$$
(2)
while for $`|y|y_0`$ the metric becomes flat:
$$A(y)\mathrm{constant}.$$
(3)
In the GRS model, this is achieved by simply patching AdS<sub>5</sub> to flat space at some point $`y=y_0`$:
$$e^{A(y)}=\{\begin{array}{cc}e^{2k|y|}\hfill & |y|y_0\hfill \\ e^{2ky_0}\hfill & |y|y_0.\hfill \end{array}$$
(4)
We can also construct examples which smoothly interpolate between AdS<sub>5</sub> and flat space :
$$e^{A(y)}=e^{2k|y|}+e^{2ky_0}.$$
(5)
In this case the cross-over from AdS to flat space occurs over a small region $`\delta yk^1`$. In both these scenarios we need $`e^{ky_0}1`$ so that the cross-over to flat space occurs at a large proper distance from the origin in the transverse space. Since the metric approaches the RS metric for $`|y|y_0`$, there is a brane located at $`y=0`$ with a positive tension $`V=6k/\kappa ^2`$ (the โPlanck braneโ) on which the matter fields will live. In the GRS model, there are additional branes at $`y=\pm y_0`$ with negative tensions $`\frac{1}{2}V`$. In the smoothed version (5), the region of negative tension is smeared over a scale $`\delta yk^1`$. As pointed out in such smooth backgrounds violate positivity, which will give rise to the instability discussed in .
In , the propagator for gravity on the positive tension Planck brane at $`y=0`$ was found to have the form
$$G(x,x^{})_{\mu \nu ,\rho \sigma }=\mathrm{\Delta }_5(x,0;x^{},0)\left(\frac{1}{2}\eta _{\mu \rho }\eta _{\nu \sigma }+\frac{1}{2}\eta _{\mu \sigma }\eta _{\nu \rho }\frac{1}{3}\eta _{\mu \nu }\eta _{\rho \sigma }\right)\frac{k}{6}\mathrm{\Delta }_4(x,x^{})\eta _{\mu \nu }\eta _{\rho \sigma }.$$
(6)
Here, the first term is what would naรฏvely be expected, since the tensor structure is five dimensional, and $`\mathrm{\Delta }_5(x,y;x^{},y^{})`$ is the scalar Greenโs function for the background (1):
$$\left(e^A\mathrm{}^{(4)}+_y^22A^{}_y\right)\mathrm{\Delta }_5(x,y;x^{},y^{})=e^{2A}\delta ^{(4)}(xx^{})\delta (yy^{}),$$
(7)
The unexpected piece is the last term in (6), which occurs because matter sources on the brane actually bend the brane, and this effect modifies the braneworld propagator. Notice that the brane-bending term involves the four-dimensional massless scalar propagator defined via
$$\mathrm{}^{(4)}\mathrm{\Delta }_4(x,x^{})=\delta ^{(4)}(xx^{}).$$
(8)
If the negative tension branes are sufficiently far away ($`e^{ky_0}1`$), then there is a large intermediate region of four-dimensional distance scales $`k^1rk^1e^{3ky_0}`$, where $`\mathrm{\Delta }_5(x,0;x^{},0)`$ is dominated by a zero energy resonance and is approximately $`k\mathrm{\Delta }_4(x,x^{})`$. In this region the two terms in (6) combine into the usual 4D graviton propagator:
$$G(x,x^{})_{\mu \nu ,\rho \sigma }kG_4(x,x^{})_{\mu \nu ,\rho \sigma }=k\mathrm{\Delta }_4(x,x^{})\left(\frac{1}{2}\eta _{\mu \rho }\eta _{\nu \sigma }+\frac{1}{2}\eta _{\mu \sigma }\eta _{\nu \rho }\frac{1}{2}\eta _{\mu \nu }\eta _{\rho \sigma }\right).$$
(9)
Hence, at these intermediate distance scales conventional gravity is recovered as a consequence of the interplay between the resonance and the bending of the brane. This makes intuitive sense: suppose we send out signals (e.g. a gravitational wave pulse) from a point on the brane, along the brane and transverse to it. If the signal along the brane only travels a distance $`rk^1e^{ky_0}`$, in the same proper time the transverse signal only explores the AdS portion of the transverse space and we expect that the physics on the brane is unchanged from the RS scenario where the AdS space extends out to infinite $`y`$ .
At larger distance scales on the brane the corresponding transverse signal reaches the flat portion of the transverse space and so the physics at scales $`rk^1e^{ky_0}`$ on the brane will be modified from pure 4D gravity. In fact there is a cross-over region at distance scales $`rk^1e^{3ky_0}`$, where $`\mathrm{\Delta }_5(x,0;,x^{},0)`$ becomes the massless scalar propagator in 5D Minkowski space. So at large distances the first term in (6) gives rise to a gravitational potential that falls off as $`1/r^2`$. However, the brane-bending term is always four dimensional and gives rise to a gravitational potential that falls off as $`1/r`$. Hence at very large distances the brane-bending term dominates and gives rise to scalar anti-gravity as noted in .
The same result has been also obtained by Pilo, Rattazzi and Zaffaroni (PRZ) , who have calculated the graviton propagator on the brane by identifying all relevant physical modes of the theory. These were shown to be the graviton resonance, and the massless radion field describing the fluctuations of the distance between the two branes, which turns out to have a negative kinetic term. The interpretation of the brane bending calculation summarized above is then that at intermediate energies, where the graviton is effectively localized, the negative kinetic term radion exactly cancels the contribution of the extra polarization in the massive graviton propagator, thus reproducing 4D Einstein gravity. For large distances however, the radion will dominate, and thus lead to the scalar antigravity as predicted in .
Below we will show how to use the technology known as the holographic renormalization group to obtain these results. In the meantime we clarify the meaning of holographic renormalization. This procedure corresponds to a renormalization group coarse graining which simply describes the effective physics on the brane. As usual, this involves integrating out degrees-of-freedom up to a certain physical length scale on the brane, which we denote as $`\stackrel{~}{r}`$, but which now also includes an averaging over the transverse space out to proper distances $`\stackrel{~}{r}`$. A proper distance $`r`$ on the brane corresponds to going out to $`y`$ in the transverse direction, where
$$r=_0^ye^{A(y^{})/2}๐y^{}.$$
(10)
This can be understood in two different ways. First, one can calculate how far into the bulk light can travel, while traveling a distance $`r`$ on the brane. For light $`ds^2=0`$, which yields $`dr=e^{A(y)/2}dy`$, thus yielding (10). Another way of obtaining the correspondence between distances on the brane and the bulk is to check how far into the bulk the horizon of an object with horizon size $`r`$ on the brane will be penetrating. This can be obtained from examining the Gregory-Laflamme instability for a black string . The result is again given by (10). Hence, physical scales $`rk^1e^{ky_0}`$ on the brane correspond to distances in the transverse space which reach out into the 5D Minkowski portion of the space in the GRS model.
Fortunately, recent developments in string theory suggest the tools we need to do the averaging which determines the effective theory on the brane at large distances. The technology is known as the holographic RG . The idea is that integrating out short distance degrees-of-freedom up to a four-dimensional length scale $`\stackrel{~}{r}`$, according to a brane observer, gives rise to an effective theory that is described by a theory with the position of the brane shifted in the transverse space to $`\stackrel{~}{y}`$, where
$$\stackrel{~}{r}=_{\mathrm{}}^{\stackrel{~}{y}}e^{A(y)/2}๐y.$$
(11)
(In the above, we are assuming that $`A(y)=2k|y|`$, the AdS form, for $`y<0`$, and so when the brane is at $`y=0`$ the cut-off is $`k^1`$.) This is illustrated in Figure 1.
More precisely, we cut out the region $`\stackrel{~}{y}y\stackrel{~}{y}`$ and re-glue the two portions as illustrated in Figure 2. The effective theory is then described by a metric
$$d\stackrel{~}{s}^2=dy^2+e^{\stackrel{~}{A}(y)}\eta _{\mu \nu }dx^\mu dx^\nu ,$$
(12)
where
$$\stackrel{~}{A}(|y|)=A(|y|+\stackrel{~}{y}),$$
(13)
and we have shifted $`|y||y|\stackrel{~}{y}`$ so that the effective brane remains at $`y=0`$. Notice that in the effective background (12) the negative tension branes of the GRS model, or the regions where they are smoothed, lie at a smaller proper distance $`k^1e^{k(y_0\stackrel{~}{y})}`$ from the effective (renormalized) brane. Notice also that for an AdS background, the effective metric (12) is identical to the original metric, expressing the fact that this situation describes a fixed-point of the RG flow.
In general, the effective brane theory will have a renormalized tension, unless the background is exactly dS<sub>5</sub>, AdS<sub>5</sub> or 5 dimensional Minkowski space. In addition to the renormalized brane tension, there could be additional operators induced in the effective brane action, encoding the effects of the metric fluctuations integrated over the strip.<sup>2</sup><sup>2</sup>2In addition to the effects we are describing there are purely quantum renormalization effects which are discussed in ref. . This is familiar from ordinary RG running:q when relevant operators are not present in the underlying (UV) theory, they can still be generated in the effective (IR) theory. The presence of these induced operators will ensure that the infrared physics is kept constant.<sup>3</sup><sup>3</sup>3The first version of this paper was missing these induced operators, and thus neglected some important effects due to fluctuations of the graviton and radion modes. First we discuss the renormalization of the effective brane tension. This is simply obtained from the Israel junction condition at the position of the effective brane:
$`\left[{\displaystyle \frac{}{y}}g_{ij}\right]|_{\stackrel{~}{y}}={\displaystyle \frac{2\kappa ^2}{3}}g_{ij}\stackrel{~}{V}(\stackrel{~}{y}),`$ (14)
which implies,
$$\stackrel{~}{V}(\stackrel{~}{y})=\frac{3A^{}(\stackrel{~}{y})}{\kappa ^2}.$$
(15)
One can obtain the same result on the effective brane tension by explicitly integrating the action over the strip between $`0`$ and $`\stackrel{~}{y}`$, and requiring that the contribution of the renormalized brane tension to the effective 4D cosmological constant matches the combined contributions of the original brane tension and the curvature of the strip we have integrated over. It is easy to see explicitly how this works in the general background metric (1). The action is given by
$`S={\displaystyle ๐yd^4x\left[\sqrt{g}\left(\frac{1}{2\kappa ^2}R+\mathrm{\Lambda }(y)\right)\right]}{\displaystyle ๐yd^4x\sqrt{\stackrel{~}{g}}\left[๐ฑ(y)+V\delta (y)\right]},`$ (16)
where $`\stackrel{~}{g}`$ is the induced four dimensional metric at constant $`y`$, and $`\mathrm{\Lambda }(y)=\frac{3}{2}\kappa ^2A^{}(y)^2`$, $`๐ฑ(y)=\frac{3}{2}\kappa ^2A^{\prime \prime }(y)`$ (away from the branes) are the static source terms needed to produce the metric (1). The source term $`๐ฑ`$ corresponds to a โsmeared brane tensionโ, not to be confused with the actual tension $`V`$ or the effective tension $`\stackrel{~}{V}(\stackrel{~}{y})`$. Note that with these conventions Einsteinโs equation is given by $`G_{ab}=\kappa ^2(\mathrm{\Lambda }(y)g_{ab}+(๐ฑ(y)+V\delta (y))g_{\mu \nu }\delta _a^\mu \delta _b^\nu )`$. Our renormalization procedure corresponds to performing the integral in the action (16) over a slice between $`0`$ and $`\stackrel{~}{y}`$, and replace this space with an effective brane with tension $`\stackrel{~}{V}(\stackrel{~}{y})`$. Thus the definition of the effective brane tension is
$$_0^{\stackrel{~}{y}}๐y\left[\sqrt{g}\left(\frac{R}{2\kappa ^2}+\mathrm{\Lambda }(y)\right)+\sqrt{\stackrel{~}{g}}๐ฑ(y)\right]+\stackrel{~}{V}(0)=\sqrt{\stackrel{~}{g}(\stackrel{~}{y})}\left(\frac{R^{(sing)}(\stackrel{~}{y})}{2\kappa ^2}\stackrel{~}{V}(\stackrel{~}{y})\right),$$
(17)
where $`R^{(sing)}(\stackrel{~}{y})`$ is the singular piece in the curvature at the position of the effective brane. Using the expression for the curvature $`R=\frac{10}{3}\mathrm{\Lambda }(y)\frac{8}{3}๐ฑ(y)\frac{8}{3}\stackrel{~}{V}(\stackrel{~}{y})\delta (y\stackrel{~}{y})`$, we find the formula for $`\stackrel{~}{V}(\stackrel{~}{y})`$ given in (15).
Let us now discuss the operators induced in the effective brane action. In a general background, there are a variety of terms. The simplest of these is a four-dimensional induced Einstein-Hilbert term of the form,
$$d^4x\sqrt{g^{ind}}R^{(4)},$$
(18)
where $`g^{ind}`$ is the induced metric at the effective brane, and $`R^{(4)}`$ is the curvature scalar calculated from the induced metric. One way to obtain the GRS model is as a limit of the two-strip model introduced in . This model is just like the GRS model with two branes, except that the second brane tension is chosen to be smaller in magnitude than $`V/2`$; therefore the space past the second brane is also $`AdS_5`$, except with a different cosmological constant. This model has a localized graviton, and is qualitatively very similar to the RS model, except that there is an additional radion mode appearing in the theory. We can calculate the effective brane action in the two-strip model, and take the limit where the second brane tension goes to $`V/2`$. Following we write the metric in the form,
$$ds^2=e^{A(y)}\left(1+B(y)f(x)\right)\left(\eta _{\mu \nu }+h_{\mu \nu }(x,y)+2ฯต(y)_\mu _\nu f(x)\right)dx^\mu dx^\nu +\left(1\frac{2B^{}(y)}{A^{}(y)}f(x)\right)dy^2.$$
(19)
Note that $`B^{}(y)/A^{}(y)`$ is a smooth function, even though $`A^{}(y)`$ and $`B^{}(y)`$ are discontinuous. Here, $`h_{\mu \nu }`$ are the four-dimensional graviton fluctuations including the zero mode of the two-strip model and the Kaluza-Klein modes, $`f`$ is the radion, and in the GRS model the background warp factor $`A(y)=2k|y|`$ for $`|y|<y_0`$ and $`A(y)=2ky_0`$ for $`|y|>y_0`$.
In order to calculate the coefficient of the induced operator (18), we will approximate the the graviton fluctuation in (19) by the zero mode $`h_{\mu \nu }(x,y)=h_{\mu \nu }(x)`$. This will be a good approximation as long as the width of the resonance is small. Expanding the 5D bulk curvature term $`\sqrt{g}R`$ to zeroth order in $`f`$ we find it contains the operator $`e^{A(y)}\sqrt{\stackrel{~}{g}}\stackrel{~}{R}^{(4)}`$, where $`\stackrel{~}{g}`$ and $`\stackrel{~}{R}^{(4)}`$ correspond to the 4D metric $`\stackrel{~}{g}_{\mu \nu }=\eta _{\mu \nu }+h_{\mu \nu }`$. Thus we can integrate this term over the strip from $`0`$ to $`\stackrel{~}{y}`$, to obtain the coefficient of the 4D Ricci scalar part of the effective brane action, $`2_0^{\stackrel{~}{y}}e^{A(y)}๐y.`$ If we wish to write this operator in the effective brane action in terms of the induced metric $`g^{ind}`$, then we have
$$S_{Ricci\mathrm{eff}}=2_0^{\stackrel{~}{y}}e^{A(y)}๐yd^4x\sqrt{\stackrel{~}{g}}\stackrel{~}{R}^{(4)}=2e^{A(\stackrel{~}{y})}_0^{\stackrel{~}{y}}e^{A(y)}๐yd^4x\sqrt{g^{ind}}R^{(4)}.$$
(20)
However, this is not the full story. In the GRS model (and the two strip model in general) the radion is relevant in the infrared, and induces additional operators in the effective brane action.
## 3 The Radion in the GRS Model
In this section we study the induced radion kinetic term on the effective brane, following the work of Pilo, Rattazzi and Zaffaroni . We start with the action
$$S=\sqrt{g}(M_{}^3R\mathrm{\Lambda }(y))d^5x\underset{i}{}\sqrt{g_i^{ind}}V_id^4x,$$
(21)
where the bulk integral can be split into two regions $`(0,y_0)`$ and $`(y_0,\mathrm{})`$ (where $`y_0`$ is the position of the negative tension brane) with cosmological constants $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$ respectively, and the brane integrals involve the brane tensions $`V_i`$ and induced metrics $`g_i^{ind}`$ on the two branes.
The form for $`B(y)`$ in the metric (19), in the region between the positive and negative tension branes is determined from the linearized Einstein equations by insisting that the 4D graviton fluctuations decouple from the radion. The four-dimensional components of the linearized Einstein equations in the bulk are proportional to,
$$e^{2ky}(2kB(y)B^{}(y))+2k(4kฯต^{}(y)+ฯต^{\prime \prime }(y))=0,$$
(22)
which is solved by ,
$$B(y)=2\left[\alpha e^{2ky}+ke^{2ky}_yฯต(y)\right].$$
(23)
The coefficient $`\alpha `$ is arbitrary, but in order for the radion coupling to matter to be normalized the same as the longitudinal graviton we choose it to be $`\alpha =1`$. In the region beyond the negative tension brane, $`y>y_0`$, $`B(y)`$ takes the form,
$$B(y)=k^{}e^{2k^{}y}_yฯต(y),$$
(24)
where for the GRS scenario we take the limit $`k^{}0.`$ In this region (for any $`k^{}`$) the radion dependent part of the metric is pure gauge and the equations of motion are satisfied for any $`f(x)`$ . The matching conditions at the positive and negative tension branes determine
$$B(0)=2,B(y_0)=2e^{2ky_0}\frac{k^{}}{(k^{}k)}.$$
(25)
Note that $`B(y_0)0`$ for the GRS model.
The pure gauge part of the action does not contribute to the effective theory on the brane, so with the above gauge choice for the metric the radion effective action is given completely by integrating over the region between the positive and negative tension brane.
As described in the previous section the fluctuation independent part of the action is made to vanish by the usual fine tuning of cosmological constants and brane tension, which for GRS is, $`\mathrm{\Lambda }_1=12k^2M_{}^3`$ between the branes, $`\mathrm{\Lambda }_2=0`$ beyond the negative tension brane; $`V_1=12kM_{}^3`$ and $`V_2=6kM_{}^3`$ are the brane tensions of the positive and negative tension brane, respectively.
We focus on the radion kinetic terms in the action. The calculation is simplified as in by expanding about the background metric, so that,
$$S_{\mathrm{radion}}=_{y_0}^{y_0}๐yd^4x\sqrt{g}\delta g^{\mu \nu }\left[M_{}^3R_{\mu \nu }\frac{1}{2}g_{\mu \nu }\left(M_{}^3R\mathrm{\Lambda }\underset{i}{}V_i\delta \left(\sqrt{g_{yy}}(yy_i)\right)\right)\right],$$
(26)
where $`\delta g^{\mu \nu }`$ is the $`๐ช\left(f(x)\right)`$ term in $`g^{\mu \nu }`$, and the equation of motion is expanded to linear order in $`f(x)`$. Given the ansatz (23) for $`B(y)`$, the equations of motion identically vanish to linear order except for the $`(yy)`$ component. This follows from the requirement that the equations of motion for $`h_{\mu \nu }`$ and $`f`$ are decoupled. The delta function contributions to the equations of motion identically vanish given the ansatz (19), which leaves a single surface term for the radion kinetic part of the action:
$$S_{\mathrm{radion}}=\frac{M_{}^3}{k}_0^{y_0}๐yd^4x\sqrt{\stackrel{~}{g}}\mathrm{\hspace{0.17em}3}B^{}(y)f\text{ }\text{ }\text{ }\text{ }\text{ }f.$$
(27)
Separating the integral over the strip $`(0,\stackrel{~}{y})`$ from the remainder of the bulk, the full action for the reduced bulk and renormalized brane is:
$`S[\stackrel{~}{y}]=`$ $`{\displaystyle _{|y|>\stackrel{~}{y}}}\sqrt{g}(M_{}^3R\mathrm{\Lambda }(y))d^5x{\displaystyle \underset{i}{}}{\displaystyle \sqrt{g_i^{ind}}V_id^4x}`$ (28)
$`+{\displaystyle d^4x\sqrt{\stackrel{~}{g}}\left[\frac{M_{}^3}{k}(1e^{2k\stackrel{~}{y}})R^{(4)}+\frac{3M_{}^3}{k}\left(B(\stackrel{~}{y})B(0)\right)f\text{ }\text{ }\text{ }\text{ }\text{ }f\right]},`$
where in the bulk integral the region $`(0,\stackrel{~}{y})`$ has been removed and the brane tension of the positive tension Planck brane $`V_1`$ is replaced by its renormalized value $`V_1(\stackrel{~}{y})`$ (which is, however, constant in the GRS scenario for $`\stackrel{~}{y}<y_0`$). When $`\stackrel{~}{y}y_0`$ there is a single tensionless brane in the effective theory, which is (aside from the ghostlike radion contribution) the DGP model . This is a new type of holographic duality between the GRS and DGP models, meaning that the DGP model is in fact (modulo the radion) nothing else but the low-energy effective theory of the GRS model.
## 4 The Graviton Propagator from Holographic RG in the GRS Model
Before we calculate the renormalized graviton Greenโs function in the GRS model, we pause to consider the rationale behind the holographic renormalization group in more detail. Performing a renormalization group transformation by moving branes through a five dimensional space may at first seem bizarre to those not familiar with recent developments in string theory. However we can arrive at a simple understanding of why this procedure is reasonable by recalling some simple classical mechanics. In order to calculate the potential due to some mass we can find an equipotential Gaussian surface and require that the integral of the gradient of the potential over this surface is equal to the mass up to appropriate factors (like Newtonโs constant). In a 5D setting we see that an equipotential surface of a gravitational source on the brane penetrates some distance into the bulk around the brane. The ratio of this distance to the corresponding distance on the brane depends in detail on the 5D metric. However as we look at longer distances on the brane, we are probing further into the bulk. Furthermore different energy/matter distributions (including distributions that penetrate into the bulk) can have the same equipotential surface. Thus we are performing an averaging over the short-distance propagation of the 5D gravitons. In GRS type geometries, at large enough distances the equipotential surface will include part of the 5D Minkowski space. At sufficiently large distances most of the interior of the equipotential surface will be 5D Minkowski and the gravitational potential will be that of 5D Minkowski space (ignoring the radion for now), falling like the inverse of the distance squared. Another way of saying this is that at distances much larger than the proper distance at which the geometry becomes flat, $`k^1e^{ky_0}`$, we would not be able to resolve the small slice of AdS<sub>5</sub> and we would effectively see a tensionless brane in 5D Minkowski space. However, this does not automatically imply that the theory on the brane will be 5D Minkowski gravity. This is because there are additional operators in the effective action on the tensionless brane, which can have important consequences. For example, the scalar anti-gravity behavior noticed in is entirely due to the radion field which is localized on the tensionless brane as envisioned in , and does not probe the bulk.
We now discuss the long distance behavior of the GRS model from the point of view of the holographic picture explained above. We have seen that at the length scale $`k^1e^{ky_0}`$ the holographic effective theory (corresponding to integrating out $`\stackrel{~}{y}`$ to that scale) is described by a tensionless brane in a 5D Minkowski space, with the action
$$d^5xM_{}^3e^{2ky_0}\sqrt{g}R+d^4x\left(M_{}^32b\sqrt{\stackrel{~}{g}}R^{(4)}cf\text{ }\text{ }\text{ }\text{ }\text{ }f\right),$$
(29)
where
$`b={\displaystyle \frac{1e^{2ky_b}}{2k}},c={\displaystyle \frac{6M_{}^3}{k}}.`$ (30)
In the first term of (29) the factor $`e^{2ky_0}`$ has been scaled out of the metric so as to restore the canonical flat 5D background metric $`g_{\mu \nu }=\eta _{\mu \nu }`$. This effective theory is nothing other than the model recently proposed by Dvali, Gabadadze and Porrati , in which there is a tensionless brane in 5D Minkowski space with an additional induced 4D curvature term, but with the ghost-like radion field necessary to cancel the effects of the extra graviton polarization automatically present on the brane. In terms of their notation the effective 5D Planck scale is given by
$`\stackrel{~}{M}^3=e^{2ky_0}M_{}^3,`$ (31)
and the effective 4D Planck scale is given by
$`M_{Pl}^2={\displaystyle \frac{M_{}^3}{k}}(1e^{2ky_0}).`$ (32)
In it was shown that the graviton propagator of this theory (without the radion) is approximately that of an almost massless 4D graviton up to distance scales of order $`\stackrel{~}{r}_0=\frac{M_{Pl}^2}{\stackrel{~}{M}^3}`$, and after that the theory becomes 5D gravity. With our parameters, $`\stackrel{~}{r}_0=k^1e^{2ky_0}`$, but since this is already the effective theory, this scale $`\stackrel{~}{r}_0`$ is related to the original scales on the brane by $`r_0=\stackrel{~}{r}_0e^{ky_0}`$, therefore we recover the results of : the GRS model reproduces ordinary 4D Einstein gravity up to the scale $`r_0=k^1e^{3ky_0}`$, and after that scale the radion will dominate and give 4D scalar antigravity (and the graviton will contribute as in ordinary 5D Minkowski gravity). This holographic approach thus sheds light on the appearance of the somewhat mysterious scale $`k^1e^{3ky_0}`$.
We now show the detailed calculation of the renormalized propagator in the GRS model. Consider renormalizing from the Planck brane to a renormalized effective brane at the position of the negative tension brane $`y_0`$. The induced brane action for the graviton on the renormalized brane is given by the second term in Eq. (29) To find the Greenโs function in the effective theory we follow the method of Giddings, Katz, and Randall . The equation for the graviton Greenโs function in transverse-traceless gauge is (after Fourier transforming from brane coordinates to the brane momentum $`p`$ and taking $`p^2=q^2`$, and setting $`M_{}=1`$ for simplicity):
$`(_y^22A^{}_y+e^Aq^2+2e^{2A}bq^2\delta (yy_0))\mathrm{\Delta }(q,y,y^{})=e^{2A}\delta (yy^{})`$ (33)
This can be put in self-adjoint form by rescaling $`\mathrm{\Delta }`$ by metric factors:
$`\mathrm{\Delta }(q,y,y^{})=e^{A(y)+A(y^{})}\widehat{\mathrm{\Delta }}(q,y,y^{}).`$ (34)
We then have
$`(_y^2+A^{\prime \prime }(A^{})^2+e^Aq^2+2e^{2A}bq^2\delta (yy_0))\widehat{\mathrm{\Delta }}=\delta (yy^{})`$ (35)
For $`y,y^{}y_0`$ the Greenโs function can be constructed by patching together the solutions of the corresponding homogeneous equation with $`y<y^{}`$ and $`y>y^{}`$, which we refer to as $`\mathrm{\Delta }_<`$ and $`\mathrm{\Delta }_>`$ respectively:
$`\widehat{\mathrm{\Delta }}=\theta (yy^{})\mathrm{\Delta }_>+\theta (y^{}y)\mathrm{\Delta }_<`$ (36)
Plugging the patched solution into Eq. (35) for $`y^{}y_0`$ yields:
$`\mathrm{\Delta }_<|_{y=y^{}}=\mathrm{\Delta }_>|_{y=y^{}}`$ (37)
$`_y(\mathrm{\Delta }_>\mathrm{\Delta }_<)|_{y=y^{}}=1`$ (38)
Since we are interested in the graviton Greenโs function and not just the scalar Greenโs function, we will impose the boundary condition implied by the Israel jump condition which relates the jump in the derivative of the metric to the brane tension. For the linearized fluctuation $`h_{ij}`$ around the background (in the absence of a Ricci scalar term in the brane action) we have
$`\left[{\displaystyle \frac{}{y}}h_{ij}\right]|_{y=y_0}=2A^{}(y_0)h_{ij}`$ (39)
Taking account of the induced Ricci scalar term in the effective brane action we find that the correct boundary condition is:
$`_y\mathrm{\Delta }_<|_{y=y_0}=(A^{}+e^{2A}b)\mathrm{\Delta }_<|_{y=y_0}.`$ (40)
In that case the implied discontinuity in the derivative of the Greenโs function at $`y=y_0`$ is just what is required by Eq. (35) due to the $`\delta `$ function piece of $`A^{\prime \prime }`$ and the induced Ricci scalar.
Setting $`y^{}=y_0`$ in Eqs. (37) and (38) and combining with Eq. (40) yields
$`_y\mathrm{\Delta }_>|_{y=y_0}+(A^{}+e^{2A}bq^2)\mathrm{\Delta }_>|_{y=\stackrel{~}{y}_0}=1`$ (41)
as the boundary condition at the renormalized brane, which we can use to determine the normalization of the homogeneous solution. The solution of the homogeneous equation in the Minkowski region are just plane waves. The choice that satisfies the outgoing wave boundary condition at infinity is
$`\mathrm{\Delta }_>=Ne^{iqye^{ky_0}},`$ (42)
where $`N`$ is a normalization factor.
Using the normalization condition (41) we can determine $`N`$ and find that the Greenโs function on the renormalized brane ($`y=y^{}=y_0`$) is given by:
$`\widehat{\mathrm{\Delta }}(q,y_0,y_0)={\displaystyle \frac{1}{iqe^{ky_0}+bq^2e^{4ky_0}}}`$ (43)
Rescaling back to the physical Greenโs function via (34) we find
$`\mathrm{\Delta }(q,y_0,y_0)={\displaystyle \frac{e^{4ky_0}}{iqe^{ky_0}+bq^2e^{4ky_0}}},`$ (44)
which agrees for $`q<ke^{ky_0}`$ with the full (and tedious) Greenโs function calculation in the two-strip model with $`k^{}0`$. For $`kqke^{3ky_0}`$, using (30) this is approximately,
$`\mathrm{\Delta }(q,y_0,y_0){\displaystyle \frac{2k}{(1e^{2ky_0})q^2}}.`$ (45)
which is the expected 4D propagator for intermediate scales. Including the 5D Planck scale couplings we can read off the effective Newtonโs constant at intermediate scales to be:
$`G_N={\displaystyle \frac{2k}{(1e^{2ky_0})M_{}^2}}`$ (46)
Alternatively, for small $`qke^{3ky_0}`$, (44) reduces to
$`\mathrm{\Delta }(q,y_0,y_0){\displaystyle \frac{ie^{3ky_0}}{q}}+{\displaystyle \frac{1}{2k}}(e^{6ky_0}e^{4ky_0}).`$ (47)
Thus we see that (neglecting the effects of the radion) at long distances the graviton Greenโs function goes over to 5D behavior, with an associated gravitational potential that falls off like $`1/r^2`$.
## 5 Cosmology of Theories with Quasi-localized Gravity
We have seen in the previous section that, as predicted in , the GRS model (and theories with quasi-localized gravity in general) interpolates between ordinary 4D Einstein gravity and (due to the presence of the radion field with negative kinetic term) 4D scalar anti-gravity. Since the cosmology of brane models has recently attracted a lot of attention , we will briefly sketch how the cosmology of the GRS model would work, even though due the radion instability this model is probably not a realistic model of the Universe. First we consider the case of pure radiation on the positive tension brane in the GRS model. Since the energy-momentum tensor for radiation is traceless, the radion does not couple to this type of source. Thus for pure radiation, the effect of the radion should be negligible at the classical level. Therefore, for a radiation dominated universe, the GRS model interpolates between the RS model at high (but below $`M_{Pl}`$) energies and a tensionless brane in 5D Minkowski space at very low energies (that is at very large distances). We expect the expansion of the Universe to be dominated by the largest scales, since the gravitational energy itself is dominated by the largest distances. This is so because the gravitational energy in a sphere grows as the square of the radius in the case of 4D gravity (and as the radius for 5D gravity). The reason is that even though the potential is decreasing with the distance, there is much more matter close to the edge of the sphere than in the middle. Therefore we will assume that the expansion of the Universe should be described by the effective holographic brane model with the cut-off given by the size of the Universe (that is, the region in causal contact since the Big Bang). This immediately indicates the type of expansion one expects here: as long as the size of the Universe is smaller than the distance where Einstein gravity turns into 4D anti-gravity, we expect the expansion to be given by that of a brane in AdS space, that is the ordinary Friedmann equation. Once the size of the Universe reaches the critical distance, it will be given by the cosmology of a tensionless brane in 5D Minkowski space. Binรฉtruy, Deffayet and Langlois (BDL) showed that the cosmology of a tensionless brane in Minkowski space does not reproduce the ordinary Friedmann equations; instead it predicts a Hubble law of the form $`H^2\rho ^2`$, where $`\rho `$ is the energy energy density of matter on the brane . Thus, once the critical distance is reached, the expansion will change to that of BDL. Since in this case $`H^2\rho ^2`$ instead of the ordinary $`H^2\rho `$, the expansion will change once the 5D phase is reached. The transition will not be as abrupt as in the largest-scale-dominance approximation, however, away from the transition region largest-scale-dominance should be a good approximation. The expansion equation for this case has been investigated in detail in . The Friedmann equation is given by
$$\frac{\ddot{a_0}}{a_0}+\left(\frac{\dot{a_0}}{a_0}\right)^2=\frac{\kappa ^4}{36}\rho (\rho +3p),$$
(48)
where $`a_0(t)`$ is the scale factor at the positive tension brane, and $`\rho ,p`$ are the radiation energy and pressure densities introduced on the brane as
$$T_\nu ^{\mathrm{brane}\mu }=b^1\delta (y)\mathrm{diag}(\rho ,p,p,p),$$
(49)
and the five dimensional metric (which includes both the background and the expansion due to the matter sources on the brane) is given by
$$ds^2=b^2(y,t)dy^2+a^2(y,t)d\stackrel{}{x}^2n^2(y,t)dt^2.$$
(50)
In it was shown, that there are two types of solutions for the above equation (48). One is the solution obtained in , for which $`a_0(t)t^{\frac{1}{4}}`$, while for the other solution
$$a_0(t)t^{\frac{1}{2}}\left(1\frac{\kappa ^4}{36}\frac{\rho ^2(t_i)a_0(t_i)^8}{t^2}+\mathrm{}\right).$$
(51)
Thus for the case of pure radiation, there are two known solutions to the expansion equations. In one solution the expansion of the Universe would slow down to $`t^{\frac{1}{4}}`$, while for the other solution (if the densities are small compared to the expansion rate at the critical distance) the solution remains an essentially unchanged $`t^{\frac{1}{2}}`$ expansion law, up to small corrections. Which solution is actually realized depends on initial conditions but we expect that for generic initial conditions (where the radiation dominated universe expands normally in the intermediate 4D regime) that the second solution should hold to good approximation.
The situation will be very different for other types of matter, for example consider an ordinary matter dominated Universe. In this case, the expansion will again start out as an ordinary 4D expansion, given by the power law $`a_0(t)t^{\frac{1}{3}}`$. However, for distances larger than the critical distance, the radion will start dominating the expansion. We will approximate the expansion equations for this case by assuming that we have a tensionless brane in 5D Minkowski space, plus the radion localized on in. At distances where the 5D nature of gravity becomes apparent, the induced 4D curvature term can be neglected compared to the 5D curvature term. In this case, one can derive the general expansion equations, which will now also involve the radion field $`\phi `$. To find the expansion equations, we start with an effective action at very large distances of the form
$$d^5x\frac{1}{2\kappa ^2}\sqrt{g}R+d^4x\frac{1}{2}_\mu \phi ^\mu \phi ,$$
(52)
where $`\phi =\sqrt{2c}f`$ represents the radion field with the wrong-signed kinetic term, and $`\kappa `$ here stands for the Newtonโs constant in the effective holographic theory, which is related by the warp factor to the fundamental parameter of the the model. Similarly to , one can derive the expansion equation for this action. The difference compared to will be that the scalar is localized on the effective brane as well, therefore they will appear as additional delta-function like sources, which will modify the โjump-equationsโ to
$$\frac{[a^{}]}{a_0b_0}=\frac{\kappa ^2}{3}\rho +\frac{\kappa ^2}{6n^2}\dot{\phi }^2,\frac{[n^{}]}{n_0b_0}=\frac{\kappa ^2}{3}(3p+2\rho )+\frac{\kappa ^2}{6n^2}\dot{\phi }^2,$$
(53)
where $`[a^{}]`$ denotes the jump of the $`y`$ derivative of the scale factor at the position of the brane. Then similarly to one obtains the expansion equation by substituting these jumps into the $`55`$ component of Einsteinโs equation. The result is given by
$$\frac{\ddot{a_0}}{a_0}+\left(\frac{\dot{a_0}}{a_0}\right)^2=\frac{\kappa ^4}{36}\rho (\rho +3p)+\frac{\kappa ^4}{72}\dot{\phi }^4+\frac{\kappa ^4}{72}\dot{\phi }^2(\rho +3p),$$
(54)
where we have rescaled time such that $`n_0=n(0)=1`$. However, this is not the whole story, since $`\rho `$ and $`p`$ are also the sources for $`\phi `$, which is given by the scalar equation of motion on the brane. The linear coupling to matter on the brane (neglecting the dimension 6 derivative couplings) is given by
$`\phi {\displaystyle \frac{\delta S}{\delta \phi }}|_{\varphi =0}=\phi {\displaystyle \frac{\delta S}{\delta \left(g^{ind}\right)^{\mu \nu }}}{\displaystyle \frac{\delta \left(g^{ind}\right)^{\mu \nu }}{\delta \phi }}=\sqrt{g^{ind}}\stackrel{~}{T}_{\mu \nu }e^{A\left(0\right)}\stackrel{~}{g}^{\mu \nu }{\displaystyle \frac{B\left(0\right)}{2}}{\displaystyle \frac{\phi \left(x\right)}{\sqrt{2c}}}.`$ (55)
Note, that the matter is assumed to be at the original Planck brane, therefore there will be no additional warp factor appearing in the coupling. Thus the radion equation of motion is
$$\text{ }\text{ }\text{ }\text{ }\text{ }\phi =\frac{1}{\sqrt{2c}}(3p\rho ),$$
(56)
which, assuming that there is only time dependence, would give the equation
$$\ddot{\phi }=\frac{1}{\sqrt{2c}}(3p\rho ).$$
(57)
Thus for general matter the expansion is determined by the coupled equations (54) and (57). Due to the antigravitational nature of the interaction mediated by the radion we expect that (like in ) the Universe would reach a maximal size and then recollapses. Since the ghostlike radion is likely to make the model unstable anyway, we will not pursue the solutions of the expansion equation sketched above.
## 6 Conclusions
We have examined a class of 5D metrics with embedded 3-branes. Along the extra dimension these models are asymptotically Minkowski, and gravitons are quasi-localized on a brane. One would expect that at sufficiently large distances the details near the brane are irrelevant and effectively there is a tensionless brane in 5D Minkowski space (a brane with non-zero tension would curve the space around it). We find that a holographic renormalization group analysis confirms this intuitive picture for the graviton. The renormalization group analysis also shows how the radion effectively gives rise to scalar โanti-gravityโ at long distances by an induced radion coupling on the brane with negative kinetic term. In fact, at intermediate and large distances the holographic effective theory is equivalent to the recently proposed model of Dvali et al., where the tensionless brane in 5D Minkowski space also has an induced 4D curvature term on the brane. Thus the behavior of quasi-localized gravity in GRS-type models at different length scales is as follows: at very short distances the theory is five dimensional (both scalar potential and tensor structure); at intermediate scales it is given by ordinary 4D gravity with corrections that can be arbitrarily small; and at ultra-large distances the graviton is again five dimensional and the 4D radion dominates. Thus these models do not seem to be internally consistent; however, if a generalized model could eliminate the radion from the light degrees of freedom (as can happen in RS models) they might produce viable cosmologies which decelerate after a late epoch. To be consistent with current observations this epoch must be later than the current epoch.
## Acknowledgements
We would like to thank Michael Graesser, Juan Maldacena, and Martin Schmaltz for useful conversations, and to Gia Dvali, Gregory Gabadadze, Massimo Porrati and Valery Rubakov for correspondence and for comments on the manuscript. We also thank Ami Katz, Luigi Pilo, Lisa Randall, Riccardo Rattazzi and Alberto Zaffaroni for comments and criticism on the first version of this paper, from which we have greatly benefitted. C.C. is an Oppenheimer Fellow at the Los Alamos National Laboratory. C.C., J.E. and T.J.H. are supported by the US Department of Energy under contract W-7405-ENG-36. J.T. is supported in part by the NSF under grant PHY-98-02709. J.T. thanks the T8 group at LANL for its hospitality while part of this work was completed. |
warning/0003/astro-ph0003220.html | ar5iv | text | # A code for optically thick and hot photoionized media
## 1 Introduction.
A number of codes were built in the past to compute the structure and the emission of photoionized media. With time, these codes became more and more sophisticated, and able to treat a larger number of situations. In the sixties they were specially designed for planetary nebulae and HII regions, i.e. for dilute and optically thin media (except in the Lyman continuum) ionized by the thermal radiation of a hot star, and were rapidly used also for ionization by a non thermal continua extending in the X-ray range, i.e. for the Narrow Line Region of Active Galactic Nuclei (AGN). At the same time another generation of codes was developed for optically thin hot collisionally ionized media, and used for the solar corona, for supernova remnants, and for the hot intergalactic gas. At the end of the seventies, observations of compact X-ray sources implying the presence of a Thomson thick medium incited Ross (1979) to develop a radiative transfer method for the continuum using a modified version of the Kompaneets equation. In the same paper he introduced the so-called โescape probabilityโ approximation to take into account diffusion and absorption into the lines. This approximation was thereafter amply used for the Broad Line Region: this was the beginning of the local escape probability era, with the fiducial paper of Kwan & Krolik (1981) aimed at studying the Broad Line Region of AGN (BLR), immediately followed by a similar computation for the atmospheres of X-ray binary stars (Kallman and McCray 1982) from which xstar derived. The most popular of these codes is cloudy, designed by Ferland (cf. for instance Ferland & Rees 1988) and continuously updated since this time (cf. Ferland 1996, Ferland et al. 1998). These codes are not only used for relatively โcoldโ clouds like the BLR, but also for warm photoionized media (Broad Absorption Line region in quasars and โWarm Absorbersโ in AGN, shock heated media, etcโฆ).
At the other extreme (i.e. for dense, non irradiated, and semi-infinite media) model atmosphere codes have been developed since the fifties with an emphasis put onto the computation of line transfer, which was treated completely, but with the LTE approximation. Non LTE effects were introduced in the sixties, and extensively studied in the seventies, with the state of the art being beautifully exposed in Mihalas book (1978). Since this time many improvments have been made in the numerical methods, in particular aiming at taking into account a large number of atoms and levels, cf. for introductory reviews Rutten 1995, Hubeny 1997.
In the nineties appeared the urgent need for โintermediateโ codes, valid for thick or semi-infinite dense media, eventually hot, irradiated by a non thermal continuum extending in the hard X-rays, to cover the whole range of situations encompassed in AGN and in binary stars. Several codes were then built, mainly to compute the em-ission spectrum of accretion disks, irradiated or not by an X-ray continuum. For irradiated disks, they were either of โphotoionization typeโ, using the Kompaneets equation (Ross & Fabian, 1993, based on the Ross 1979 code), or coupling an existing photoionization code to a Monte Carlo computation (Zycki et al. 1994), to take into account Compton diffusions. On the other hand, sophisticated model atmosphere codes were transformed to be applied to accretion disk structure and emission (Hubeny, 1990, Hubeny & Hubeny 1997 and 1998), but without the external irradiation by an X-ray continuum.
Owing to the high optical thickness of the medium in several frequency ranges, such codes require that the transfer of both the continuum and the lines be solved in an โexactโ way, i.e. avoiding approximations such as local escape approximation (for the lines) or one stream approximation (for the continuum). Since the medium is generally dense and sometimes close to LTE, they require that all processes and inverse processes be carefully handled. Being irradiated by an X-ray continuum, the medium contains a large number of ionic species, from low to high ionization, which should all be introduced in the computation. Finally, the medium being hot and thick, not only Thomson, but also Compton scattering, should be taken into account.
We have undertaken to build a code in order to satisfy these requirements. Precisely we have built several interconnected codes, which allow more flexibility. The ensemble is far from being perfect and still contains several approximations which restrict its use, but we intend to improve it in the future.
We have presently four codes. One of them (titan) is designed to study the structure of a warm or hot thick photoionized gas, and to compute its emission - reflection - transmission spectrum from the infrared up to about 20 KeV. It solves the energy balance, the ionization and the statistical equilibria, the transfer equations, in a plane-parallel geometry, for the lines and continuum. Then, given the thermal and ionization stratification, the computation of the emitted spectrum from 1 KeV to a few hundre-ds KeV is performed with noar which uses a Monte-Carlo method taking into account direct and inverse Compton scattering (it allows also to study various geometries). pegas adresses the case of more diluted and thin media, and is similar to cloudy with however some differences, and iris is specially devoted to the computation of the line fluxes, including very weak lines, using the most recent available atomic data and a full treatment of the statistical equilibrium equations for a great number of levels. These codes can be used to model a wide variety of astrophysical media, particularly those in which the energy transport is purely radiative, but also plasmas heated by other mechanisms (viscosity, shocks, energetic particulesโฆ). They have already been used in several published papers (Collin-Souffrin et al. 1996, Czerny & Dumont 1998, Porquet et al. 1998, Abrassart 1999).
The three codes noar, pegas and iris, are described in other papers (Abrassart 2000, Dumont & Porquet 2000). In this paper we present titan and we discuss a few simple cases, schematized by a slab of gas irradiated on one side or on both sides. As far as possible we will perf-orm comparisons with cloudy in the range of parameters where both codes can be used, to assess the validity of titan. However detailed comparisons are not possible, as they use quite different transfer methods in particular. We will also show some examples of the capabilities of titan in the parameter range which is not accessible to cloudy.
titan is mainly designed to determine the structure (temperature and ionization state) and the continuum em-ission spectrum of a thick hot photoionized slab of gas. Owing to the representation of the ions made in the code it does not give an accurate determination of the weak lines. For a detailed determination of the line spectrum of some ions, it should be complemented by iris.
We briefly summarize below the physical processes (Sect. 2), the transfer method and the iteration procedure (Sect. 3), the thermal equilibrium, (Sect. 4), focussing only on the aspects which are not treated in a standard way. The coupling of titan and noar is briefly described in Sect. 5, and some applications are presented in Sect. 6.
## 2 Physical processes
In titan the physical state of the gas (temperature, ion abundances and level populations of all ionic species) is computed at each depth, assuming stationary state, i.e. local balance between ionizations and recombinations of ions, excitations and deexcitations, local energy balance (equality of heating due to absorption and cooling due to local emission), and finally total energy balance (equality between inward and outward fluxes).
Due to the large range of density and temperature inside the medium, many physical processes play a role at some place and should therefore be taken into account. Ionization equilibrium equations include radiative ionizations by continuum and line photons, collisional ionizations and recombinations, radiative and dielectronic recombinations, charge transfer by H and He atoms, the Auger effects, and ionizations by high energy electrons arising from ionizations by X-ray photons. Energy balance equations include free-free, free-bound and line cooling, and Compton heating/cooling.
The emission-absorption mechanisms for the continuum include free-free and free-bound processes, two-photon process, and Thomson scattering. Special care is given to recombinations to ground state, which are very important from an observational point of view in the X-ray range. They are treated differently according to the relative values of $`kT`$ to the photon energy bin, in order to get an accurate frequency dependence.
Hydrogen and hydrogen-like ions are treated as 6-level atoms. Levels 2s and 2p are treated separately, while full l-mixing is assumed for higher levels. All processes including collisional and radiative ionizations and recombinations are taken into account for each level (cf. Mihalas 1978). Recombinations onto levels $`n>5`$ are not taken into account, which amounts to assuming that the higher levels are in LTE with the continuum, which is generally true in the conditions for which this code is presently used. In the future we plane to add several other levels, and to sum the contributions of the higher levels as it is done for instance in cloudy. Level populations are then obtained as usual by matrix inversion.
In order to save computation time, non H-like ions are presently treated with a rough approximation: interlocking between excited levels is neglected and populations of the excited levels are computed separately using a two-level approximation. This approximation does not predict correctly the details of the line spectrum, since it neglects subordinate lines. Nevertheless recombinations onto excited states are taken into account in the ionization equilibrium and the transfer of these photons is treated in an approximative way as proposed by Canfield & Ricchiazzi (1980). We assume also that each recombination produces a resonant photon after cascades.
The gas composition include 10 elements (H, He, C, N, O, Ne, Mg, Si, S, Fe), and all their ionic species are taken into account. Photoionization cross sections are fitted from Reilman & Manson (1979) and Band L.M. et al. (1990), these values being correct as far as neutral and once ionized ions are not concerned, and it is the case in hot media. For total radiative and dielectronic recombination rates, we use Aldrovandi & Pรฉquignotโs data (1973). When possible, collisional excitation rates are taken from the Daresbury Report (1985). Most of data for iron come from Arnaud & Raymond (1992), Kaastra & Mewe (1993), and from Fuhr et al. (1988). Ionizations by high energy electrons arising from ionizations by X-range photons are taken from Bergeron & Souffrin (1973). Inverse processes (except dielectronic recombinations) are computed through the equations of detailed balance. In the case of a gas close to LTE, we neglect dielectronic recombinations for consistency to insure the balance for each process. All induced processes are taken into account.
The equations are not recalled as they have been given in previously quoted papers.
Iron K lines
These lines require special attention as they are intense in Seyfert nuclei, and they will be observed in detail in the future with Chandra and XMM. Though the iron K lines constitute a complex system described in Band D.L. et al. (1990), we assume presently only one โmeanโ line per ion with an oscillator strength equal to 0.4, as suggested by Band. More detailed computations are not required as far as the Doppler broadening of the lines (or Compton broadening as well, see Abrassart 2000) is much larger than the distance between the lines, as it is the case in AGN.
In other computations of the line fluxes it is generally assumed that resonant trapping of K$`\alpha `$ photons of FeXVII to Fe XXIII (Ross & Fabian 1993) suppress completely these lines when the Thomson thickness of the emitting medium is larger than a given value, of the order of 0.02 (Zycki & Czerny 1994). Here the transfer of these lines is handled in a standard way, with an additional term included in the statistical and ionization equations to take into account the competition between radiative deexcitation and the Auger process. The population $`N_i^k`$ of the upper level of a transition K$`\alpha `$ of the ion $`i`$ is given by:
$$N_i^kA_{ki}=y_{i1}(N_{i1}K_{i1}+N_iB_{ik}J_{ik})$$
(1)
while the ionization equation for the same ion $`i`$ writes:
$`N_i[(P_i`$ $`+`$ $`K_i)+B_{ik}J_{ik}(1y_{i1})]`$
$`=`$ $`N_{i+1}\alpha _iN_{i1}K_{i1}(1y_{i1}).`$
$`A_{ki}`$ and $`B_{ik}`$ are the Einstein coefficients of the K$`\alpha `$ line, $`J_{ik}`$ is the mean intensity integrated over the line profile, $`y_i`$ is the fluorescent yield, $`\alpha _i`$ is the recombination coefficient, and $`P_i`$ (respt. $`K_i`$) is the photoionization rate of the ground state (respt. of the K-shell) of the ion $`i`$. The second term on the left side of Eq. 2 is generally only of the order of $`10\%`$ of the total photoionization rate.
## 3 Radiation transfer
An important quantity will be used all along this paper, the illumination (or โionizationโ) parameter. Among several definitions used in the literature, we adopt the following:
$$\xi =\frac{4\pi F_{\mathrm{inc}}}{n_\mathrm{H}},$$
(3)
where $`F_{\mathrm{inc}}`$ is the frequency integrated flux incident on one side of the slab and $`n_\mathrm{H}`$ is the hydrogen number density at this surface. Note that it does not preclude the possibility of having also a flux incident on the back side of the slab. We call the attention on the fact that all along this paper we will mimic a semi-isotropic and not a mono-directional illumination.
### 3.1 Present method for the transfer of the continuum
We want to study, on one hand media with an optical thickness larger than unity in a range of frequency rich in emission or in absorption lines, on the other hand media with a very inhomogeneous structure. For instance if the illuminated side has a temperature of a few 10<sup>6</sup> K, while the temperature is only a few 10<sup>4</sup> K on the back side of the cloud, the spectrum emitted by the illuminated side is mainly formed in hot layers where the absorption coefficient is weak at all wavelengths and Thomson diffusion is dominant (we shall call it the โreflected spectrumโ though it does not correspond to pure reflection). On the contrary the spectrum emitted by the backside of the slab is formed in cold layers where absorption must be carefully taken into account (we shall call it the โoutward spectrumโ). Note also that, even if the incident illumination is mono-directional, the โtransmitted spectrumโ is almost isotropic since the Thomson thickness, $`\tau _\mathrm{T}`$, is larger than unity. It is therefore not differentiated from the โoutward spectrumโ even when the incident source is not located on the line of sight. It is possible to estimate an approximate transmitted spectrum using the expression $`F_\nu ^{\mathrm{transm}}=F_\nu ^{\mathrm{inc}}exp(\tau _\nu )`$, where $`F_\nu ^{\mathrm{inc}}`$ is the incident flux, in an aim of comparison.
The transfer is treated with the Eddington two-stream approximation, i.e. the intensity is assumed to be constant in each hemisphere. The transfer equations can then be written :
$`{\displaystyle \frac{1}{\sqrt{3}}}{\displaystyle \frac{dI_\nu ^+}{dz}}`$ $`=`$ $`(\kappa _\nu +{\displaystyle \frac{\sigma }{2}})I_\nu ^++{\displaystyle \frac{\sigma }{2}}I_\nu ^{}+ฯต_\nu `$ (4)
$`{\displaystyle \frac{1}{\sqrt{3}}}{\displaystyle \frac{dI_\nu ^{}}{dz}}`$ $`=`$ $`(\kappa _\nu +{\displaystyle \frac{\sigma }{2}})I_\nu ^{}+{\displaystyle \frac{\sigma }{2}}I_\nu ^++ฯต_\nu `$
where $`z`$ is the distance to the illuminated edge, $`\kappa _\nu `$ is the absorption coefficient, $`\sigma `$ is the diffusion coefficient - here it is due to Thomson scattering - and $`ฯต_\nu `$ is the emissivity (all these quantities are local and depend on the frequency $`\nu `$ except $`\sigma `$). Note that this approximation is closer to the semi-isotropic case than to the normal case. We have already mentioned that it is appropriate for the reflected and for the transmitted flux when the Thomson thickness is larger than unity. It is also appropriate for a semi-isotropic illumination, if the source of radiation is extended (like for instance in the disk-corona model of AGN, or in the blob model of Collin-Souffrin et al. 1996).
The mean intensity $`J_\nu `$ is equal to $`(I_\nu ^++I_\nu ^{})/2`$, while the flux F defined as usual by $`I_\nu \mathrm{cos}\theta d\omega `$ is equal to $`(I_\nu ^+I_\nu ^{})2\pi /\sqrt{3}`$ ; so that the reflected flux is equal to $`I_\nu ^{}(0)2\pi /\sqrt{3}`$, and the outward flux to $`I_\nu ^+(H)2\pi /\sqrt{3}`$. The optical depth and the total optical depth are defined as :
$$\tau _\nu (z)=_0^z\sqrt{3}(\kappa _\nu +\sigma )๐z^{}\mathrm{and}T_\nu =\tau _\nu (H)$$
(5)
where $`H`$ is the total thickness of the slab, and the source function as:
$$S=\frac{(ฯต_\nu +\sigma J_\nu )}{(\kappa _\nu +\sigma )}$$
(6)
Both sides of the slab can be illuminated by an external radiation, so the boundary conditions are, at $`z=0`$:
$$I_\nu ^+(0)=\frac{\sqrt{3}}{2\pi }F_\nu ^{\mathrm{inc}},$$
(7)
and at $`z=H`$:
$$I_\nu ^{}(H)=\frac{\sqrt{3}}{2\pi }F_\nu ^{\mathrm{back}}$$
(8)
where $`F_\nu ^{\mathrm{back}}`$ is the back-side flux (it is equal to zero in many of the following computations).
The formal solution of the transfer equations between $`z\delta z`$ and $`z`$ gives:
$$I_\nu ^+(z)=I_\nu ^+(z\delta z)e^{\delta \tau _\nu }+e^{\tau _\nu }_{\tau _\nu \delta \tau _\nu }^{\tau _\nu }S_\nu (t)e^{+t}๐t.$$
(9)
Assuming that $`S_\nu (t)`$ is nearly constant in the interval $`z,z+\delta z`$ , one gets:
$$I_\nu ^+(z)=I_\nu ^+(z\delta z)e^{\delta \tau _\nu }+\frac{1e^{\delta \tau _\nu }}{2}[S_\nu (z\delta z)+S_\nu (z)]$$
(10)
with a similar equation for $`I_\nu ^{}(z)`$.
Guessing the initial value of $`I_\nu ^{}(0)`$ fails because calculations diverge, unless this initial value is provided with an extreme precision: the predicted value should differ from the real one at all frequencies by less than $`10^6`$. This problem has also been pointed out by Coleman (1993). We therefore start to compute $`I_\nu ^{}(z)`$ from the back side with:
$$I_\nu ^{}(z)=I_\nu ^{}(z+\delta z)e^{\delta \tau _\nu }+\frac{1e^{\delta \tau _\nu }}{2}[S_\nu (z+\delta z)+S_\nu (z)]$$
(11)
o Thus we calculate $`S_\nu (z)`$ and the mean intensity $`J_\nu (z)`$ from Eq. 6 and:
$`J_\nu (z)`$ $`=`$ $`{\displaystyle \frac{I_\nu ^+(z\delta z)+I_\nu ^{}(z+\delta z)}{2}}e^{\delta \tau _\nu }`$
$`+`$ $`{\displaystyle \frac{(1e^{\delta \tau _\nu })}{4}}[S_\nu (z\delta z)+2S_\nu (z)+S_\nu (z+\delta z)]`$
assuming a constant $`\delta z`$ (actually it is not constant so the formulae are more complicated but do not deserve to be given here).
The procedure is then the following. We divide the sl-ab into a set of plane-parallel layers with the density $`n_\mathrm{H}`$ given in each layer. Note that the code can easily be coupled with a prescription for the density or for the pressure, such as a constant pressure, or a pressure given by hydrostatic equilibrium. Presently the slab is divided in about 300 layers. The geometrical thickness of each layer is sm-aller close to the surfaces than in the middle of the slab because of the rapid variation of the physical paramet-ers (the optical thickness, the temperature, the ionization state).
\- For each layer, starting from the illuminated side, where $`I_\nu ^+(0)`$ is given by Eq. 7, the ionization and thermal balance equations are solved. The source function, the opacity and the emissivity, are computed. As they depend on $`J_\nu `$, we use Eq. 3.1 with the value of $`I_\nu ^+(z\delta z)`$ of the previous layer and the value of $`I_\nu ^{}(z+\delta z)`$ provided by the previous iteration and given by Eq. 11 (see below).
\- $`I_\nu ^+(z)`$ is transferred through the layer according to Eq. 10, while $`S_\nu (z)`$ and $`\tau _\nu (z)`$ are buffered for each layer and each frequency.
\- when the back side of the cloud is reached, new values of $`I_\nu ^{}(z)`$ are calculated, starting from the back side where $`I_\nu ^{}(H)`$ is given by Eq. 8, and using Eq. 11 with the previous values of $`S_\nu (z)`$ and $`\tau _\nu (z)`$ ; actually we use the values given by several previous iterations to accelerate the procedure.
\- the whole calculation is repeated until convergence. It is stopped when energy balance is achieved for the whole slab (i.e. when the flux entering on both sides of the slab is equal to the flux coming out from both sides).
The drawback of this method is that it requires a large number of iterations to get a complete convergence of the model. The iteration starts by assuming given values for $`I_\nu ^{}(z)`$. If this initialisation is not adequate, convergence problems are encountered. We have chosen for the first iteration $`I_\nu ^{}(z)`$ equal to $`I_\nu ^+(z)`$ multiplied by a factor of the order of unity, which works reasonably well. Actually the convergence of $`I_\nu ^{}(z)`$ (as well as of the temperature and the ionization state) is reached within less than 1$`\%`$ after only a relatively small number of iterations in the layers corresponding to about one Thomson thickness. This property is interesting as far as the outward spectrum is negligible, which is the case in the EUV and X-ray range. It can also be used to determine the ionization state and the temperature of the layers where Compton diffusion takes place (cf. Sect. 5 about coupling with the code noar).
Fig. 1 displays the reflected continuum obtained after different numbers of iterations (for clarity we prefer to show here the continuum and not the complete spectrum). The model is a slab with the following characteristics:
\- a constant density 10<sup>12</sup> cm<sup>-3</sup>,
\- a column density 10<sup>26</sup> cm<sup>-2</sup>;
\- it is illuminated on one side by an incident power law continuum proportional to $`\nu ^1`$ from 0.1 eV to 100 KeV,
\- the illumination parameter is $`\xi =10^3`$ erg cm s<sup>-1</sup>;
\- there is no illumination on the other side;
\- the abundances are (in number): H: 1, He: 0.085, C: 3.3 10<sup>-4</sup>, N: 9.1 10<sup>-5</sup>, O: 6.6 10<sup>-4</sup>, Ne: 8.3 10<sup>-5</sup>, Mg: 2.6 10<sup>-5</sup>, Si: 3.3 10<sup>-5</sup>, S: 1.6 10<sup>-5</sup>, Fe: 3.2 10<sup>-5</sup>.
In the following we will call this our โreference modelโ. All along the paper we will use the same density and the same spectral distribution of the illuminating radiation, so we will not specify these parameters anymore. Note that, for this peculiar spectral distribution, $`\xi `$ and the ratio of incident ionizing photon number to hydrogen densities, U, are related by : $`\xi =8.21U\mathrm{ln}(\nu _{\mathrm{max}}/\nu _{\mathrm{min}})`$ where $`\nu _{\mathrm{max}}`$ and $`\nu _{\mathrm{min}}`$ are the maximum and minimum photon frequencies.
We want to stress here the importance of the backward intensity. It cannot be neglected as it participates to the ionization, to the level population, and to the energy balance. For instance in a purely scattering medium the returning flux varies as $`F_{\mathrm{inc}}\frac{\tau _\mathrm{T}/2}{1+\tau _\mathrm{T}/2}`$, so when $`\tau _\mathrm{T}`$ is much larger than unity, all the incident flux is reflected. Thus the temperature of the first layer at the illuminated side (which depends on the sum of incident plus returning radiation) should increase with the column density for a given incident flux (except possibly in cases of very high values of the illumination parameter).
Fig. 2 shows the dependence of the surface temperature on the column density $`N`$. The model is similar to the reference one except that the illumination parameter is equal to 100 erg cm s<sup>-1</sup> and the column density is varying. We see that the surface temperature is constant up to about $`N=10^{22}`$ cm<sup>-2</sup>, but after it increases from 2 10<sup>5</sup> K to 2.7 10<sup>5</sup> K for $`N`$ increasing from $`10^{22}`$ to $`10^{24}`$ cm<sup>-2</sup>. It saturates when a large fraction of the slab becomes close to LTE, and the returning flux becomes constant. With cloudy (version 9004) the surface temperature does not depend on the column density (for a correct comparison the illumination parameter in cloudy must be multiplied by a factor $`\sqrt{3}`$ to account for the semi-isotropic approximation used in titan).
Fig. 3 displays the dependence of the surface temperature on an assigned ratio $`I^{}(0)/I^+(0)`$ independent of the frequency. We see that a modest returning flux (20$`\%`$) is sufficient to account for the 60$`\%`$ increase of the surface temperature reached for $`N=10^{24}`$ cm<sup>-2</sup>, with respect to a very small column density. To understand this behaviour, one should remember that the spectral distribution of the reflected flux contains a much larger proportion of EUV radiation than the incident continuum, in particular in the lines. So the illumination parameter is increased by a larger amount than the ratio of the integrated reflected over incident continuum. We can also guess that the increase of the temperature at the illuminated surface with respect to cloudy will be compensated by a decrease at the back side, because of energy conservation (cf. Fig. 14).
### 3.2 Present method for line transfer
Radiative transfer of the lines is treated in the same way as the continuum, the line profile being represented by several frequencies distributed symmetrically around the line center.
We assume that even if large macroscopic velocities are present, they do not play any role in the line transfer, in other words that the thickness of the slab is much smaller than the scale length of the velocity gradient. We also assume that the lines do not overlap, which is valid as far as multiplets are treated globally, and if the microscopic turbulent velocity is not much larger than the thermal velocity (doing this we neglect line fluorescence due to the overlap of two lines, like the HeII Balmer$`\beta `$ line and the hydrogen Ly$`\alpha `$ line, whose wavelengths differ only by a fraction 4 10<sup>-4</sup>).
Finally we assume presently complete redistribution in all the lines. We intend to include partial redistribution for resonance lines like Ly$`\alpha `$ in the near future. For the moment partial redistribution in these lines can be mimicked by assuming complete redistribution in a pure Doppler profile, as the line photons diffuse more in space than in frequency, owing to the absence of enlargment of the lower level. A Doppler profile can also be used for strongly interlocked lines, and in any case, it does not differ strongly from a Voigt profile when $`\tau _0`$ is smaller than $`1/a`$ so the medium is optically thin in the damping wings. Note that partial redistribution is less important in dense media, like irradiated accretion disks in AGN, than in dilute media, owing to collisional redistribution.
The line profile $`\varphi _\nu `$ is a symmetrical Voigt function. The Doppler width $`\mathrm{\Delta }\nu _\mathrm{d}`$ and the damping constant depend on the temperature and on the density, and consequently vary with $`z`$. It may happen that the temperature decreases by two orders of magnitude from the illuminated to the back side, so $`\mathrm{\Delta }\nu _\mathrm{d}`$ decreases by one order of magnitude. This must be taken into account to choose the points in the profile. To solve the statistical equilibrium equations one needs to know the total line intensity weighted by the profile $`J_\nu \varphi (\nu )๐\nu `$, and to treat the thermal and ionization equilibria one simply integrates the line intensity over the frequencies $`J_\nu ๐\nu `$. For intense lines the first integral is dominated by the Doppler core, and the second by the wings. In the case of a saturated line with strong wings, the set of frequencies chosen to describe the line must cover a range from a few tenths $`\mathrm{\Delta }\nu _\mathrm{d}`$ to a few hundreds $`\mathrm{\Delta }\nu _\mathrm{d}`$ and at least 10 points are necessary for a correct representation of the profile.
Fig. 4 displays for the reference model the variation with $`z/H`$ of the OVIII Ly$`\alpha `$ profile in both directions, $`I_\nu ^{}`$ and $`I_\nu ^+`$, $`J_\nu `$ being the half sum of these profiles. We see that the profile varies considerably from the surface to the deeper layers.
Now we are able to show the difference between the transfer treatment and the escape probability approximation for the lines discussed in the Appendix. In this aim, we have compared the divergence flux $`\rho `$ ( Eq. 17) with the escape probability towards the illuminated side $`P_e(\tau _0)`$, and with an approximate expression proposed by Rees et al. (1989) to take into account continuum absorption in the statistical and in the energy balance equations :
$$\beta _{\mathrm{eff}}=\frac{\kappa _c}{\kappa _c+\kappa _l}+\frac{\kappa _l}{\kappa _c+\kappa _l}\frac{[P_e(\tau _0)+P_e(T_0\tau _0)]}{2}$$
(13)
where $`\beta _{\mathrm{eff}}`$ is the total effective escape probability, $`\kappa _l`$ and $`\kappa _c`$ are the absorption coefficients respectively at the line center and in the underlying continuum, $`T_0`$ is the total optical depth of the slab at the line center, and $`P_e(\tau _0)`$ (respt. $`P_e(T_0\tau _0)`$) are the escape probabilities towards the illuminated (respt. the back) side.
$`P_e(\tau _0)/2`$ and $`\beta _{\mathrm{eff}}`$ have been computed with titan and compared to $`\rho `$ in Fig. 5, as functions of the distance $`z`$ to the illuminated surface, for the reference model.
We see on this figure that $`\beta _{\mathrm{eff}}`$ and $`P_e(\tau _0)/2`$ are almost always identical. It is first due to the fact that the slab is very thick, so $`P_e(\tau _0)`$ is much larger than $`P_e(T_0\tau _0)`$ except very close to the back side, and second because $`\kappa _c`$ is always $`\kappa _l`$, as all the lines displayed on the figure are intense. But $`\beta _{\mathrm{eff}}`$ differs considerably from $`\rho `$ in the region where the lines are formed (i.e. the region where $`\rho `$ is not equal to zero and where the ion is present). In particular $`\rho `$ is negative in an important fraction of the line formation region for H$`\alpha `$, L$`\alpha `$, and OVIII L$`\alpha `$, while $`\beta _{\mathrm{eff}}`$ is a positive quantity.
As a consequence the line spectrum is not correctly computed when escape probabilities are used. This can be seen on Table 1 which gives the equivalent widths (negative if the line is in absorption, positive if it is in emission) of some intense lines from both sides of the slab. Several of these lines are in absorption, either from the illuminated or from the dark side, which is impossible with the escape probability approximation.
Presently the radiative transfer is solved for 300 frequencies in the continuum from 0.01 eV to 25 keV, including all the ionization edges, and for 440 spectral lines including the fluorescence Fe K lines, the integrals over the line profiles being achieved using 15-point Gauss-Legendre quadrature, together with a change of variable.
Remarks on the convergence of the method
It is well-known that the lambda iteration method used here (actually mainly for historical reasons, because we st-arted to build a nebular code with the line escape probability approximation) converges extremely slowly for intense lines in stellar atmospheres, and we intend to replace it by the Accelerated Lambda Iteration method as soon as possible. The convergence of $`J_\nu `$ is achieved after 600 iterations, for all lines except a few ones and for all continua. This can be checked on Fig. 6 which displays for the same lines as Fig. 5 the profiles of the lines emitted by the bright side, after different number of iterations, for the reference model. As a consequence, the structure of the slab (T, ionization) does not change at all after a few hundreds iterations. Indeed our case is more favorable than stellar atmospheres. The explanation is double: 1. the relatively small density of the medium compared to a stellar atmosphere, which decreases comparatively the influence of collisional excitations, and 2. the fact that in a photoionized plasma the temperature is low for a high ionization degree, so the excited levels lie at very high energies compared to the thermal energy. Consequently the populations of these highly ionized ions are smaller than at Boltzman equilibrium with the ground level, and the lack of convergence for excited level populations does not have a strong influence on the ionization state of the corresponding ion. We have also checked that the same structure is obtained for different initial boundary conditions. All this implies that the continuum emitted from both sides (including for instance the Lyman discontinuity) is completely converged, and only the few non converged lines are not well computed, in particular those emitted by the back side (but they are negligible in the whole energy balance). Since we do not claim to compute a correct line spectrum before the subordinate lines are implemented in the code, we think that this approximation is sufficient for our present purpose.
## 4 Energy balance
titan has been designed for stationary radiatively heated media, so the energy conservation implies the balance of radiative flux. However it can be used also for other cases out of radiative equilibrium, for instance by imposing the temperature of the medium, or by adding a non radiative energy in each layer, such as the viscous flux of an accretion disk. We describe here how the code works in a purely radiative case.
We first stress the fact that the energy balance should be realized both locally and globally.
The local energy balance equation writes:
$$\frac{dF_\nu }{dz}๐\nu =0=n_en_H[\mathrm{\Gamma }_{\mathrm{tot}}+\mathrm{\Lambda }_{\mathrm{tot}}],$$
(14)
where $`n_en_H\mathrm{\Gamma }_{\mathrm{tot}}`$ and $`n_en_H\mathrm{\Lambda }_{\mathrm{tot}}`$ are the total heating and cooling rates per unit volume. The temperature of a layer is computed through an iteration process until $`\mathrm{\Gamma }_{\mathrm{tot}}`$ and $`\mathrm{\Lambda }_{\mathrm{tot}}`$ are equal.
As an example, Fig. 7 gives the dependence of the cooling and heating rates, and of the equilibrium temperature, on the distance from the illuminated side $`z`$, for the reference model. We note that in this model Compton heating-cooling is not very important, and that bound-free contributes as a heating term, compensated by a cooling term due to lines and free-free processes. Note also that a quasi LTE is reached at a few Thomson depths in this model (it is reached at a smaller optical depth for a sm-aller value of $`\xi `$). Above this value of $`z`$ the temperature is almost constant, except close to the back side where it decreases more rapidly due to the leakage of photons.
The global energy balance is reached through an iteration procedure, when the total flux emitted by both sides of the slab, $`F_{\mathrm{tot}}^{\mathrm{em}}`$, is equal to the total flux incident on both sides, $`F_{\mathrm{tot}}^{\mathrm{inc}}`$. Actually this is the most important convergence criterium and it requires many iterations. To illustrate the problem, Fig. 8 displays the flux of a few intense emission lines and of the total flux in the lines and continuum, for the reference model, as functions of the number of iterations. We see that after about 200 iterations the global energy balance is reached within about one percent.
In the same way, Fig.9 shows that the energy balance with titan is reached within one percent after 200 iterations for two models with a column density equal to 10<sup>25</sup> cm<sup>-2</sup>, and with an illumination parameter equal to $`\xi =100`$ and $`\xi =1000`$ erg cm s<sup>-1</sup>. cloudy or xstar have been used several times in this range of parameters (Martocchia & Matt, 1996, Zycki et al. 1994, for instance), but it is not sure that the global energy balance is achieved owing to the escape treatment of the lines, as discussed above.
Finally we must recall that the neglect of subordinate lines (except in hydrogen-like ions) could lead to a substantial increase of the energy losses and consequently to a change of the thermal balance, as these lines have a smaller optical thickness than resonant lines, and therefore can escape more easily from the deep layers. Since in a photoionized plasma the temperature is relatively low and the excited level populations of highly ionized ions are generally small, one could expect only a small influence on the overall spectrum.
To check this point we compare in Figs. 10 and 11 the temperature and the emitted spectrum for the reference model, with and without taking into account interlocking between excited levels in hydrogen like ions (note that interlocking is included for hydrogen and helium in both cases). These ions are the most abundant ones in the region emitting the reflected spectrum, in our reference model. Fig. 10 shows that the temperature is slightly shifted in layers corresponding to a Thomson thickness of a few. As a consequence the emitted spectrum is also different, but not by an important amount. We can infer a similar behaviour for smaller illumination parameters, even if the abundant ions are less ionized, as the temperature is also lower. For a higher illumination parameter the lines contribute less to the energy balance.
Since the detailed computed line spectrum of hydrogen-like ions is different with the two treatments, contrary to the overall spectrum, it is mandatory to treat correctly the excited levels for all the ions in order to get a correct line spectrum. Note that in this respect cloudy is presently doing much better than titan, as it includes subordinate lines and more levels for each ion.
Compton heating and cooling
The Compton heating-cooling term is equal to:
$$\mathrm{\Gamma }\mathrm{\Lambda }=\frac{\sigma _T}{m_ec^2}\frac{1}{n_H}4k(T_{\mathrm{Comp}}T)4\pi J_\nu ๐\nu $$
(15)
where $`T_{\mathrm{Comp}}`$ is the Compton temperature. We give here this well-known equation only to recall that it depends (as well as $`T_{\mathrm{Comp}}`$) on the spectral distribution of the local flux, so it requires a careful computation of this flux (including the lines).
Below 25 KeV, $`J_\nu `$ is provided directly by titan in each layer. Above 25 KeV, $`J_\nu `$ is not calculated by titan, but by the code noar, as the energy shift of the photons due to Compton scattering is no more negligible. In this case the results from noar are also used to compute the exact value of the Compton heating-cooling term. This is discussed in more detail in the next section.
## 5 Coupling with the Monte Carlo code
In order to compute the effect of Compton scattering on the high energy part of the reprocessed radiation, and particularly on the reflected spectrum above 1 keV, and to determine precisely the Compton heating-cooling rate entering in the energy balance equation, we need another numerical approach. The transfer scheme in titan is based on frequency arrays, and inelastic scattering would introduce a redistribution on energy bins making the code much heavier. In this aim Abrassart (2000) developped a Monte Carlo code, noar. The asset of such an approach is also that it enables to investigate an arbitrary geometry and to determine the angular dependance of the observed spectra. These aspects were difficult to tackle with a code such as titan. Moreover, it allows to easily extract time variability information.
noar is described in detail in Abrassart (2000). We explain here only how it is coupled with titan. It shares the same opacity data and includes the same set of 102 ions, for the 10 elements of highest cosmic abundances. Given all the ionisation degrees provided by titan, it computes absorption cross sections in each layer. Free-free absorption is also taken into account. noar includes the recombination continua of hydrogen and helium like ions, but does not include line emission, except fluorescence lines. The proper yields to account for the competition with the Auger effect are used for these lines. Fluorescence of Fe XVII-XXIII is suppressed by resonnant trapping. For the pseudo fluorescence of H and He-like species, a simplifying prescription is adopted which includes only the two most probable outcomes of a K shell photoionization, i.e. direct recombination on ground level or L-K transition. The spectral distribution of the recombination continuum is determined by the local temperature in the โcurrentโ slab.
noar takes into account direct and inverse Compton scattering. The method used for modelling Compton scattering is basically the one described in Pozdniakov, Sobol & Sunyaev (1983) and in Gorecki & Wilczewski (1984), with a different use of statistical weight, because of the competition with photoelectric absorption. When the high energy cut-off of the incident continuum is of the order of 100 KeV, the reflected spectrum above 10 KeV exhibits a โCompton humpโ peaking at about 30 KeV (Lightmann & White 1988). This spectral region of optimal continuum albedo occurs where the sum of photoelectric absorption and Compton losses is minimum. This feature depends weakly on the ionization state, although its prominence in an observed X-ray spectra, and notably its level with respect to the iron line, is sensitive to $`\xi `$ (see Fig. 19) as the soft X-ray albedo. Note that the shape and extension of the hump depends on the high energy cut-off of the incident spectrum, which is 100 KeV all along this paper.
Fluorescent lines are significantly Compton broadened for rather high illumination parameter, above $`\xi 10^3`$ erg cm s<sup>-1</sup>. The broadening is asymmetyric, the profile is skewed toward the red. At still higher $`\xi `$ and temperature, a blue wing appears, due to the inverse Compton effe-ct. The extend and importance of both wings depends on the optical depth of the scattering medium. The red wing depends only on the frequency of the line (i.e. on the ionization state of iron) whereas the blue one only depends on the temperature of the scattering medium.
titan and noar lead to almost identical reflected spectra in the 1-20 KeV range (they differ only by Compton broadening of lines and of photoelectric edges), and noar allows to obtain the spectrum in the higher energy range. This spectrum is used for the global energy balance.
Another use of noar is to provide titan with the local Compton gains and losses in each layer. This is necessary, because Compton heating-cooling is dominated by energy losses of photons $`>`$ 25 Kev, in particular for high values of $`\xi `$ when Compton heating-cooling is dominating the energy balance and should be computed in an exact way.
As an example Fig. 12 gives the dependence of the ratio $`(\mathrm{\Gamma }_{\mathrm{Comp}}\mathrm{\Lambda }_{\mathrm{Comp}})/\xi `$ on the Thomson depth, for different values of $`\xi `$, the other parameters being the same as in the reference model. We see that the ratio does not depend on the illumination parameter if $`\xi 10^4`$ erg cm s<sup>-1</sup>. This is expected since the Compton heating-cooling is dominated by high energy photons, which do not care about the ionization state. It also means that inverse Compton is negligible as a cooling process (though it has an influence on the line spectrum), otherwise $`\frac{\mathrm{\Gamma }\mathrm{\Lambda }}{\xi }`$ would depend on the gas temperature and therefore on $`\xi `$. This is due to the relatively low temperature of the gas compared to the mean energy of the photons. This property breaks for very high illumination parameters, since low energy photons are then able to penetrate in the deepest layers and to play a role in the cooling rate, and the temperature of the gas is close to $`T_{\mathrm{Comp}}`$.
The deposition of energy shown on Fig. 12 corresp-onds to a semi-isotropic distribution of the incident radiation. It is different in the case of a mono-directional radiation. This is shown on Fig. 13 which displays $`(\mathrm{\Gamma }_{\mathrm{Comp}}\mathrm{\Lambda }_{\mathrm{Comp}})/\xi `$ for a neutral gas in the case of an isotropic and of a perpendicular illumination. When the radiation field is semi-istropic there is a rapid decrease of $`\frac{\mathrm{\Gamma }\mathrm{\Lambda }}{\xi }`$ for small values of $`\tau _\mathrm{T}`$, due to the small distance covered by the photons. This is compensated by a smaller (by about a factor two) deposition of energy in the deeper layers.
The Compton heating-cooling rate obtained with noar is then fitted analytically as a function of $`z`$, which is transferred to titan.
## 6 Some applications
A full grid of models will be published elsewhere (Dumont & Abrassart 2000). Here we discuss only a few interesting cases.
### 6.1 Optically thin hot medium: the Warm Absorber in AGN
The Warm Absorber (WA) is a hot medium located on the line of sight of the X-ray source in AGN. It is responsible of the absorption edges of OVII and OVIII observed in their soft X-ray spectrum in about 50$`\%`$ of them. It might also constitute the โmirrorโ invoked in the Unified Scheme to account for the polarized Broad Lines observed in Seyfert 2 galaxies (Antonnucci & Miller 1985). It has been modelled in many papers, using ion (Netzer 1993, 1996), xstar (Krolik & Kriss 1995), cloudy (Nicastro, Fiore & Matt 1999 for the most recent paper), pegas and iris (Porquet et al. 1999). This medium has a column density of 10<sup>21</sup> to 10<sup>24</sup> cm<sup>-2</sup>, and a typical illumination parameter of $`\xi =100`$ erg cm s<sup>-1</sup>. Porquet et al. have shown that its density should be at least 10<sup>10</sup> cm<sup>-3</sup> to avoid the emission of too strong coronal lines.
As the WA is relatively thin ($`\tau _\nu 1`$ for the continuum in a large range of frequencies), it is possible to use the line escape probability formalism. This is however limited to the case where $`\tau _\nu `$ is smaller than unity in the continuum underlying intense lines. This specific problem will be discussed in detail in the paper describing the code pegas (Dumont & Porquet 2000). A correct computation of the backward flux is also required. We have shown indeed that the temperature depends on the intensity of this backward flux.
We have run several models corresponding to the conditions of the WA, i.e. for $`\xi =100`$ erg cm s<sup>-1</sup> and column densities varying from 10<sup>22</sup> to 10<sup>24</sup> cm<sup>-2</sup> (the other parameters being the same as in the reference model). We have compared the results to those obtained with cloudy for the same models, to check that they are similar at low column densities, and to determine at which column densities they begin to differ strongly. Note that a detailed comparison between titan and cloudy is not possible, as these codes differ both in the transfer method and in the atomic data (for instance owing to the semi-isotropic illumination, a given column density corresponds to a larger optical thickness with titan than with cloudy), so only trends can be obtained.
Fig. 14 displays the dependence of the temperature on $`z`$, obtained with titan and with cloudy. As discussed previously, it is higher with titan at the illuminated side for a high column density, owing to the returning flux, and it is lower at the dark side.
The ionization state (of fundamental importance for the study of the WA) and the emission spectrum, are also different with titan and with cloudy at high column densities. Fig. 15 shows the ionization state of Carbon, Oxygen and Iron for a column density of 10<sup>24</sup> cm<sup>-2</sup>, computed with titan and with cloudy. While the ionization state close to the illuminated side is quite similar (in particular if one remembers that there is a factor $`\sqrt{3}`$ in the optical depth), it is very different in the deepest layers, where cloudy leads to a smaller degree of ionization than titan (CIII and OIII instead of CIV and OIV). This effect cannot be explained by the difference in optical thickness, which acts in the opposite direction. It is certainly due to the different treatments of the diffuse continuum, and perhaps of the lines. With titan we have checked that the diffuse continuum at large depths is sufficient to ionize OIII into OIV. To illustrate this point Fig. 16 displays the variation of the flux at 54.9 eV (the OIII-OIV ionization edge) as a function of the depth. It shows that the diffuse flux is still very intense and sufficient to ionize OIII at large depths (contrary to the transmitted flux which is very weak). Note that this result depends strongly on the way the spectral distribution of the diffuse flux is computed. If it would be reduced to a unique frequency at the recombination frequency of He<sup>++</sup>, 54.4 eV, (as it is generally the case with the โOn The Spotโ approximation), there will be no photons able to ionize OIII.
Fig. 17 displays the reflected and outward spectra, for column densities 10<sup>23</sup> and 10<sup>24</sup> cm<sup>-2</sup>. We see that the overall shape of the continuum is very similar with titan and with cloudy. However, some differences in the detailed features appear for the largest column density. For instance the Lyman edge is in emission in the reflected spectrum with both codes, but it is larger with titan (presumably because the temperature is higher in the em-ission layers). In the outward spectrum it is present as a very weak absorption with cloudy, while it is in emission with titan.
The line spectra are obviously also different. There are much more lines with cloudy, but on the other hand the fewer lines of titan are more intense. This is expected since subordinate lines are not taken into account exce-pt for hydrogen-like ions in titan. As mentioned before, some lines appear in absorption with titan, both in the reflected and in the outward spectrum.
### 6.2 Hot medium optically thick to Compton scattering: the UV-soft X-ray emitting medium in AGN
titan is specially designed for Compton thick photoionized media like those commonly assumed to emit the UV/soft X-ray continuum in AGN and to produce through Compton reflection the FeK $``$7 KeV line and the 30 KeV hump observed in many Seyfert 1 galaxies.
Although the exact nature of this medium is not known (an irradiated accretion disc, as proposed by Ross & Fabian 1993, Zycki et al. 1994, to quote only the first papers on the subject, or a clumpy Compton thick medium, cf. Collin-Souffrin et al. 1996), its characteristics and physical state are comparable: a shell of gas with a column density of at least 10<sup>25</sup> cm<sup>-2</sup>, a temperature of 10<sup>5</sup>-10<sup>6</sup> K due to radiative heating, and a density spanning a range from 10<sup>12</sup> to 10<sup>15</sup> cm<sup>-3</sup>. The corresponding value of $`\xi `$ is typically 300 to 3000 erg cm s<sup>-1</sup>. Our reference model is therefore representative of this medium. In the case of an irradiated accretion disk, it has been mimicked as a sh-ell of constant density irradiated from above by a power law continuum, and from below by a black body radiation (actually it is not the best way to take into account the presence of the underlying viscously heated disk, cf. Rozanska et al. 2000).
Let us consider first a slab illuminated on one side by a semi-isotropic radiation. Fig. 18 displays $`T`$ versus $`z`$, for different values of $`\xi `$. The other parameters are the same as in the reference model. It is interesting to notice that the quasi constant temperature regime (corresponding to LTE) is reached for increasing values of $`z`$ when $`\xi `$ increases. For instance it is reached for $`z10^{13}`$ cm, corresponding to $`\tau _\mathrm{T}6`$ for $`\xi =1000`$ erg cm s<sup>-1</sup>. It means that any computation aiming at giving the reflected spectrum of such a model should solve correctly the thermal and ionization equilibrium until these deep layers.
Fig. 19 shows the reflected spectrum for the same models. Note that the resolution adopted here (100) does not correspond to a region where the lines are broadened by large velocity field, either turbulent or organized (rotation). We use this high resolution only to show more clearly the details of the spectrum.
We see that the overall shape of the spectrum computed with titan and noar is very similar in the range 1-20 KeV. However the detailed spectral features are different, particularly for the higher values of $`\xi `$. For instance the complex feature near 7 KeV (which is made of a mixture of several Iron edges and of Iron lines dominated by low ionization stages for $`\xi =100`$ erg cm s<sup>-1</sup>, and by FeXXV and FeXXVI for $`\xi =10^4)`$ erg cm s<sup>-1</sup> is strongly modified by Compton scattering. For $`\xi =10^4`$ erg cm s<sup>-1</sup> the line has a large red wing due to direct Compton scattering, and a weak blue wing due to inverse Compton scattering. The absorption edge is also completely erased. The effect is smaller for $`\xi =10^3`$ erg cm s<sup>-1</sup>, and almost absent for $`\xi =100`$ erg cm s<sup>-1</sup>, as discussed above. Similar results have been obtained in many papers dealing with the X-ray spectrum of Seyfert galaxies (cf. for instance Ross et al. 1999).
Let us now compare the results obtained by us with titan, and by Zycki et al. (1994), for a slab of constant density 10<sup>14</sup> cm<sup>-3</sup>, illuminated by a power law $`\nu ^{0.9}`$ between 10 eV and 100 KeV on one side, and by a black body $`T_{\mathrm{BB}}=10^5`$K on the other side. We will compare the results for $`\xi `$(Zycki et al.)=300, which corresponds to $`\xi `$(titan)$``$ 200 erg cm s<sup>-1</sup>, owing to our semi-isotropic illumination, and to a slightly different definition of $`\xi `$.
Like us, Zycki et al. used a Monte-Carlo method to take into account Compton scattering. The thermal and ionization state of the gas is calculated apart with the code xstar, using like cloudy the escape probability approximation. Fig. 20 gives $`T`$ versus $`\tau _\mathrm{T}`$ obtained by us and by Zicky et al. The same result is observed as with cloudy: our temperature is larger near the surface of the slab. It reaches $`T_{\mathrm{BB}}`$ at $`\tau _\mathrm{T}`$=0.7. For a larger value of $`\xi `$, the connection with the underlying black body would be reached in even deeper layers, as discussed above. For $`\tau _\mathrm{T}0.7`$ the temperature becomes smaller than $`T_{\mathrm{BB}}`$, and joins $`T_{\mathrm{BB}}`$ only at $`\tau _\mathrm{T}20`$. This behaviour is however dependent on the thickness of the slab, it is why this representation of the disk is not meaningful (Rozanska et al. 2000). On the contrary Zycki et al. impose the gas temperature to be equal to $`min(T_{\mathrm{BB}},T_{\mathrm{thermal}\mathrm{balance}})`$.
Fig. 21 displays the reflected spectrum, computed by us and by Zycki et al. The resolution is degraded to 20, to fit the energy bins of Zycki et al. The overall shape of the spectra are roughly similar. The major discrepancy is the OVIII edge, in emission with our computation and in absorption with Zycki et al.โs one. We think that it is due to the higher surface temperature and to the lower temperature in the deeper layers obtained in our computation.
## 7 Conclusion
We have shown that the coupling of titan and noar allows to compute the structure and the emission of hot Compton thick irradiated media in an unprecendented way. First it solves consistently both the global and the local energy balance, which is impossible in a thick medium with codes handling the line transfer with the escape probability approximation, as all present photoionization codes do. We have also shown the importance of the returning flux (which is neglected in photoionization codes) even for relatively low column densities. Second it takes into account in an exact way inverse and direct Compton scattering, both in the energy balance and in the computation of the emitted spectrum. Finally, it allows to treat any geometry, open or closed.
Although the problem of the convergence process is not as drastic as in stellar atmospheres, since we are able to get complete converged structure and spectrum in a still reasonable computing time, a most urgent improvement of the code is to accelerate the convergence process through the use of the Accelerated Lambda Iteration method. This will not only allow to get the results in a much smaller time, but also to get convergence for the few lines which are still not converged after about 10<sup>3</sup> iterations.
Then the following improvements of titan will be to take into account subordinate lines in solving a multi-level atom for all ions and to bring up to date the atomic data. More detailed line spectra for some abundant ions will be obtained through coupling with the code iris. The L shells of Iron which are already taken into account for FeXVII to FeXXII will be implemented as well as a better representation of the Iron K lines. A few elements will be added to the already ten existing ones.
###### Acknowledgements.
We are grateful to the referee, G. Ferland, for his comments which have led to substantially clarify the paper, to M-C. Artru for a very careful reading of the manuscript, and to I. Hubeny for enlightning discussions.
## Appendix : Previous methods
### 7.1 Transfer of the continuum
It is not always clear which approximations are used in the different codes. Basically the computation of the diffuse radiation field seems to be similar in all codes, except in Ross & Fabian (1993) and in subsequent works using this code, where it is computed using the Kompaneets equation. cloudy (Ferland 1996), xstar (Kallman & Krolik, 1995 for the last version), ion (Netzer 1990, 1993), use a modified version of the โon the spotโ approximation, which amounts to assuming a kind of escape probability for the diffuse continuum, and one stream approximation in the transfer equation. These approximations are correct only for a continuum optical thickness smaller than unity (effective or total opacity), and when the properties of the cloud do not vary considerably between the point where a photon is emitted and the point where it is reabsorbed. In particular one very important requirement for studying hot and thick media is to take into account the radiation returning from the backside of the slab, even when it is not illuminated. As the medium is optically thick in a large frequency range, this radiation is intense and modifies the physical state of the whole slab, including the illuminated side(see Fig. 2).
### 7.2 Line transfer: the escape probability approximation
All previous photoionization codes, except that of Collin-Souffrin and Dumont (S.) (1986), treat line transfer by the so-called local escape probability formalism, which uses the probability that a line photon emitted at a given point can escape in a single flight from the cloud, assuming that the rest of the cloud is homogeneous and has the same properties as the emitting layer. It amounts to identifying the divergence flux of the statistical equilibrium equations with the escape probability intervening in the line emerging flux. This approximation is valid only in the case of complete frequency redistribution, absence of line interlocking, and if the medium is homogeneous. Its use can have severe consequences on the emission line spectrum and on the energy balance when these conditions are not fulfilled as it was often discussed, see for example by Collin & S.Dumont (1986) or Elitzur (1984). We introduce here the divergence flux in the aim to show that our results are completely different from those obtained by the escape probability formalism.
To simplify the discussion, let us consider in the rest of this section a simple two-level atom. The level population balance writes:
$`n_u(A_{ul}+B_{ul}{\displaystyle J_\nu \psi _\nu ๐\nu }+n_\mathrm{e}C_{ul})`$ (16)
$`=n_l(B_{lu}{\displaystyle J_\nu \varphi _\nu ๐\nu }+n_\mathrm{e}C_{lu})`$
where $`\varphi _\nu `$ and $`\psi _\nu `$ are the absorption and emission line profiles, $`A_{ul}`$, $`B_{ul}`$ and $`B_{lu}`$, $`C_{ul}`$ and $`C_{lu}`$ are the usual radiative (Einstein) and collisional excitation and deexcitation coefficients, and $`J_\nu `$ is the angle averaged intensity.
Let us now define the divergence flux of the transition, $`\rho _{ul}`$:
$$\rho _{ul}=\frac{\left\{n_u(A_{ul}+B_{ul}J_\nu \psi _\nu ๐\nu )n_l(B_{lu}J_\nu \varphi _\nu ๐\nu )\right\}}{n_uA_{ul}}.$$
(17)
With this definition Eq. 7.2 becomes simply:
$$n_u(n_\mathrm{e}C_{ul}+\rho _{ul}A_{ul})=n_ln_\mathrm{e}C_{lu}.$$
(18)
Let us assume the line is the only contributor to radiation. The energy emitted locally in the line in a small volume of surface 1 cm<sup>-2</sup> and length $`dz`$ writes:
$`dF`$ $`=`$ $`h\nu [n_u(A_{ul}+B_{ul}{\displaystyle J_\nu \psi _\nu ๐\nu })n_lB_{lu}{\displaystyle J_\nu \varphi _\nu ๐\nu }]dz`$ (19)
$`=`$ $`\rho _{ul}n_uAulh\nu dz=\rho _{ul}4\pi ฯตdz`$
where $`ฯต`$ is the emissivity coefficient $`n_uA_{ul}h\nu /4\pi `$. Integrating this equation over depth, one gets the emerging line flux:
$$F=4\pi \rho _{ul}\epsilon ๐z$$
(20)
According to this equation the divergence flux seems to be equal to the probability that, once emitted at a distance $`z`$ from the surface, a photon can escape from the medium, which is called the โescape probabilityโ $`P_e`$.
From this fact, if one makes the approximation that the emission and the absorption profiles are equal, one computes this escape probability by integrating the attenuation over the line profile:
$$P_e\frac{1}{2}\frac{_0^1\frac{h\nu }{4\pi }n_uA_{ul}\varphi _\nu exp(\tau \varphi _\nu /\mu )๐\nu ๐\mu }{\frac{h\nu }{4\pi }n_uA_{ul}\varphi _\nu ๐\nu }$$
(21)
where $`\mu `$ = cos$`\theta `$, and it gives the usual expressions for complete redistribution in a Voigt profile:
$$P_e(\tau _0)max(\frac{1}{1+2\tau _0\sqrt{\pi ln(\tau _0+1)}},\frac{2}{3}\sqrt{\frac{a}{\tau _0\sqrt{\pi }}}),$$
(22)
where $`\tau _0`$ is the optical thickness at the line center and $`a`$ is the usual damping constant. This expression can thus replace $`\rho _{ul}`$ in Eqs. 18 and 20.
With this method the frequency dependent transfer eq-uations are successfully replaced by frequency integrated ones, and solved consistently with the statistical equat-ions.
Several estimates of the escape probability are used in the literature (see for example Kwan & Krolik, 1981, Rees, Netzer & Ferland, 1989, Kallman & McCray, 1982, Ko & Kallman, 1994), to account for partial redistribution, for line or continuum interlocking, and for continuum absorption. For instance several ways to take into account the continuum opacity (or the overlapping of two lines) have been proposed (Elitzur & Netzer, 1984, Netzer, Elitzur & Ferland, 1985). In the case of the Broad Line Region of AGN, which is optically thin for the continuum underlying the main lines, Collin-Souffrin et al. (1981), Avrett & Loeser (1988), have discussed the influence of different approximations for partial or complete redistribution, and how they compare to an exact transfer treatment, and Collin-Souffrin & Dumont S. (1986) have shown that the escape probability approximation is roughly valid (it can however lead to a decrease by a factor as large as two of the ratio of a lower to a higher transition, such as H$`\alpha `$/H$`\beta `$ or L$`\alpha `$/H$`\alpha `$).
However these approximations are not valid if $`\tau _{\mathrm{cont}}`$ is of the order of unity at the line frequency and if the line photons are created in one place and absorbed in another place where the physical conditions are different. This happens for instance for EUV and soft X-ray lines created in the hot region and absorbed in ionizing He<sup>+</sup> ions in a cooler region.
If the local escape probability approximation breaks down, it has a severe consequence: the emitted line spectrum and the thermal and the ionization balance are not correctly computed. For instance the divergence flux takes frequently negative values in our computations, even for intense lines, and even in the reflected spectrum. This is not allowed with escape probabilities. We illustrate this discussion with some examples in Sect. 3.2. |
warning/0003/hep-ph0003014.html | ar5iv | text | # The decay ๐โข(782)โ5โข๐ in chiral theory.
## Abstract
The arguments are put forward that the many pion decays $`\omega 2\pi ^+2\pi ^{}\pi ^0`$ and $`\pi ^+\pi ^{}3\pi ^0`$ provide an ideal test site for testing the predictions of chiral models of the vector meson decays into many pions. Using the approach based on the Weinberg Lagrangian or, in a new language, the Lagrangian of hidden local symmetry added with the term induced by the anomalous Lagrangian of Wess and Zumino, the partial widths of these decays are evaluated, and their excitation curves in $`e^+e^{}`$ annihilation are obtained. The discussed are the perspectives of the experimental study of the decays $`\omega 5\pi `$ in $`e^+e^{}`$ annihilation and photoproduction.
At present, the unconventional, from the point of view of the chiral pion dynamics, sources of soft pions are feasible. Indeed, the progress in increasing the luminosity of low energy $`e^+e^{}`$ colliders ($`\varphi `$ factories) could offer the naturally controlled sources of soft pions. Other possible source of such pions could be the intense photon beams , provided the sufficiently low invariant mass regions of the many pion systems are isolated. The yield of pions is considerably enhanced when they are produced through proper vector resonance states. Then, choosing the many pion decays of sufficiently low lying resonances, one can obtain the soft pions in quantities sufficient for testing the predictions of chiral models that include vector mesons.
The decay $`\omega (782)5\pi `$ whose final state pions possess the momenta $`|๐ช_\pi |74`$ MeV, is just of this kind. The latter value is sufficiently small to expect the manifestation of chiral dynamics in most clean form. By this we mean that the higher derivative and loop terms in the effective Lagrangian are severely suppressed. The present paper is devoted to the evaluation of the partial width of this decay and plotting its excitation curve in $`e^+e^{}`$ annihilation.
The $`\rho \pi `$ sector is considered here on the basis of the Weinberg Lagrangian revived later as the Lagrangian of hidden local symmetry (HLS) . The former looks as
$``$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(_\mu ๐_\nu _\nu ๐_\mu +g[๐_\mu \times ๐_\nu ]\right)^2`$ (3)
$`+{\displaystyle \frac{m_\rho ^2}{2}}\left[๐_\mu +{\displaystyle \frac{[๐
\times _\mu ๐
]}{2gf_\pi ^2(1+๐
^2/4f_\pi ^2)}}\right]^2`$
$`+{\displaystyle \frac{(_\mu ๐
)^2}{2\left(1+๐
^2/4f_\pi ^2\right)^2}}{\displaystyle \frac{m_\pi ^2๐
^2}{2(1+๐
^2/4f_\pi ^2)}},`$
where $`๐
,m_\pi `$ and $`๐,m_\rho `$ stand for the isovector field and mass of pion and $`\rho `$ meson, respectively, cross denotes the vector product in the isovector space, and $`f_\pi =92.4`$ MeV is the pion decay constant. The $`\rho \rho \rho `$ coupling constant $`g`$ and the $`\rho \pi \pi `$ coupling constant $`g_{\rho \pi \pi }`$ are related to the $`\rho `$ mass and pion decay constant $`f_\pi `$ via the parameter of hidden local symmetry $`a`$ as
$`g`$ $`=`$ $`m_\rho /f_\pi \sqrt{a},`$ (4)
$`g_{\rho \pi \pi }`$ $`=`$ $`\sqrt{a}m_\rho /2f_\pi .`$ (5)
Note that $`a=2`$, if one demands the universality condition $`g=g_{\rho \pi \pi }`$ to be satisfied. Then the so called Kawarabayashi-Suzuki-Riazzuddin- Fayyazuddin (KSRF) relation arises
$$2g_{\rho \pi \pi }^2f_\pi ^2/m_\rho ^2=1$$
(6)
which beautifully agrees with experiment. The $`\rho \pi \pi `$ coupling constant resulting from this relation is $`g_{\rho \pi \pi }=5.89`$. The inclusion of the interaction of the $`\omega (782)`$ with the $`\rho \pi `$ state is achieved upon adding the term induced by the anomalous Lagrangian of Wess and Zumino ,
$$_{\omega \rho \pi }=\frac{N_cg^2}{8\pi ^2f_\pi }\epsilon _{\mu \nu \lambda \sigma }_\mu \omega _\nu \left(๐
_\lambda ๐_\sigma \right),$$
(7)
where $`N_c=3`$ is the number of colors, $`\omega _\nu `$ is the field of $`\omega `$ meson.
One may convince oneself that the $`\omega \rho \pi 5\pi `$ decay amplitude unambiguously results from the Weinberg Lagrangian Eq. (3) and the anomaly induced Lagrangian (7). This amplitude is represented by the diagrams shown in Fig. 1. As one can foresee, its general expression looks cumbersome. However, it can be considerably simplified upon noting that the small pion momenta permit one to use the nonrelativistic expressions,
$`M(\rho ^0\pi _{q_1}^+\pi _{q_2}^+\pi _{q_3}^{}\pi _{q_4}^{})`$ $``$ $`{\displaystyle \frac{g_{\rho \pi \pi }}{2f_\pi ^2}}(\epsilon ,q_1+q_2q_3q_4),`$ (8)
$`M(\rho ^0\pi _{q_1}^+\pi _{q_2}^{}\pi _{q_3}^0\pi _{q_4}^0)`$ $``$ $`{\displaystyle \frac{g_{\rho \pi \pi }}{4f_\pi ^2}}(\epsilon ,q_1q_2),`$ (9)
$`M(\rho ^+\pi _{q_1}^+\pi _{q_2}^+\pi _{q_3}^{}\pi _{q_4}^0)`$ $``$ $`{\displaystyle \frac{g_{\rho \pi \pi }}{4f_\pi ^2}}(\epsilon ,q_1+q_22q_4),`$ (10)
$`M(\rho ^+\pi _{q_1}^+\pi _{q_2}^0\pi _{q_3}^0\pi _{q_4}^0)`$ $``$ $`{\displaystyle \frac{g_{\rho \pi \pi }}{f_\pi ^2}}(\epsilon ,q_1),`$ (11)
for the $`\rho 4\pi `$ decay amplitudes in the diagrams Fig. 1(a), where $`\epsilon `$ stands for the $`\rho `$ meson polarization four-vector The notation for the Lorentz invariant product of two four-vectors $`a`$ and $`b`$ is $`(a,b)=a_0b_o\mathrm{๐๐}`$. . They are valid with the accuracy $`5\%`$ in the $`4\pi `$ mass range relevant for the present purpose. Likewise, the expression for the combination $`D_\pi ^1M(\pi 3\pi )`$ standing in the expression for the diagrams in Fig. 1(b) can be replaced, with the same accuracy, by $`(8m_\pi ^2)^1`$ times the nonrelativistic $`\pi 3\pi `$ amplitudes. The latter look as
$`M(\pi ^+\pi ^+\pi ^+\pi ^{})`$ $`=`$ $`2m_\pi ^2/f_\pi ^2,`$ (12)
$`M(\pi ^+\pi ^+\pi ^0\pi ^0)`$ $`=`$ $`m_\pi ^2/f_\pi ^2,`$ (13)
$`M(\pi ^0\pi ^+\pi ^{}\pi ^0)`$ $`=`$ $`m_\pi ^2/f_\pi ^2,`$ (14)
$`M(\pi ^0\pi ^0\pi ^0\pi ^0)`$ $`=`$ $`3m_\pi ^2/f_\pi ^2.`$ (15)
Note that, in the nonrelativistic limit, the $`\rho 4\pi `$ decay amplitudes depend on the HLS parameter $`a`$ only through an overall factor $`g_{\rho \pi \pi }/f_\pi ^2=\sqrt{a}m_\rho /2f_\pi ^3,`$ while the $`\pi 3\pi `$ amplitudes do not depend on it at all.
The LHS approach permits one to include the axial mesons as well <sup>ยง</sup><sup>ยง</sup>ยงThe problem of inclusion of vector, axial mesons, and photons in the framework of chiral theories has demanded considerable efforts. See . It is solved in an elegant way in the HLS approach .. An ideal treatment would consist of that under the assumption of $`m_\rho Em_{a_1}`$, the difference between the models with and without $`a_1`$ meson would be reduced to the allowing for the higher derivatives and would be small Allowing for the higher derivatives demands also allowing for the chiral loops, the task which is not yet fulfilled for vector mesons.. In particular, the correction to the $`\rho ^04\pi `$ decay width due to $`a_1`$ meson is estimated at the level of 30-40 $`\%`$ . However, since the invariant mass of four pions in the diagrams Fig. 1(a) is less than 642 MeV, the $`a_1`$ contribution in the $`\omega 5\pi `$ decay is severely suppressed. Furthermore, the introducing $`a_1`$ meson results in the terms $`(_\mu ๐
)^4`$ in the Lagrangian, but their contribution to the $`\pi 3\pi `$ transition amplitude is inessential in the relevant three pion invariant mass region less than 500 MeV.
Note also that the maximum pion momentum $`|๐ช_\pi |`$ is about 115 MeV, but the contribution of such regions of the phase space is negligible. The dominant contribution comes from the momentum $`|๐ช_\pi |70`$ MeV. In the kinematical situation of the $`\omega 5\pi `$ decay we also neglect the contributions of other higher derivatives and quantum corrections due to chiral loops. Thus, our approach is zeroth order approximation to the $`\omega 5\pi `$ decay amplitude in the framework of the Lagrangians (3) and (7), in a close analogy with the Weinberg amplitudes in classical $`\pi \pi `$ scattering. In this approximation all chiral models of the vector meson interactions with pions indistinguishable, and any difference could manifest upon including the higher derivatives and chiral loops.
Then one obtains, upon neglecting the corrections of the order of $`O(|๐ช_\pi |^4/m_\pi ^4)`$ or higher, the expression for decay amplitudes:
$`M(\omega 2\pi ^+2\pi ^{}\pi ^0)`$ $`=`$ $`{\displaystyle \frac{N_cg_{\rho \pi \pi }g^2}{8(2\pi )^2f_\pi ^3}}\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu \{(1+P_{12})q_{1\lambda }[{\displaystyle \frac{(q_2+3q_4)_\sigma }{D_\rho (qq_1)}}{\displaystyle \frac{2q_{4\sigma }}{D_\rho (q_1+q_4)}}]`$ (18)
$`(1+P_{35})q_{3\lambda }\left[{\displaystyle \frac{(q_5+3q_4)_\sigma }{D_\rho (qq_3)}}{\displaystyle \frac{2q_{4\sigma }}{D_\rho (q_3+q_4)}}\right]`$
$`(1+P_{12})(1+P_{35})q_{3\lambda }[{\displaystyle \frac{2q_{4_\sigma }}{D_\rho (qq_4)}}+{\displaystyle \frac{q_{1\sigma }}{D_\rho (q_1+q_3)}}]\},`$
with the final momentum assignment according to $`\pi ^+(q_1)\pi ^+(q_2)\pi ^{}(q_3)\pi ^{}(q_5)\pi ^0(q_4)`$, and
$`M(\omega \pi ^+\pi ^{}3\pi ^0)`$ $`=`$ $`{\displaystyle \frac{N_cg_{\rho \pi \pi }g^2}{8(2\pi )^2f_\pi ^3}}(1P_{12})(1+P_{34}+P_{35})\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu q_{1\lambda }`$ (21)
$`\times \{q_{3\sigma }[{\displaystyle \frac{1}{D_\rho (qq_3)}}{\displaystyle \frac{1}{D_\rho (q_1+q_3)}}]`$
$`q_{2\sigma }[{\displaystyle \frac{4}{3D_\rho (qq_1)}}{\displaystyle \frac{1}{2D_\rho (q_1+q_2)}}]\},`$
with the final momentum assignment according to $`\pi ^+(q_1)\pi ^{}(q_2)\pi ^0(q_3)\pi ^0(q_4)\pi ^0(q_5)`$. Note that the first term in each square bracket refers to the specific diagram shown in Fig. 1(a) while the second one does to the diagram shown in Fig. 1(b). In both above formulas, $`ฯต_\nu `$, $`q_\mu `$ stand for four-vectors of polarization and momentum of $`\omega `$ meson, $`P_{ij}`$ is the operator of the interchange the pion momenta $`q_i`$ and $`q_j`$, and
$$D_\rho (q)=m_\rho ^2q^2i\sqrt{q^2}\mathrm{\Gamma }_{\rho \pi \pi }(q^2)$$
(22)
is the inverse propagator of $`\rho `$ meson whose width is dominated by the $`\pi \pi `$ decay mode:
$$\mathrm{\Gamma }_{\rho \pi \pi }(q^2)=\mathrm{\Gamma }_{\rho \pi \pi }(m_\rho ^2)\frac{m_\rho ^2}{q^2}\left(\frac{q^24m_\pi ^2}{m_\rho ^24m_\pi ^2}\right)^{3/2}.$$
(23)
Yet even in this simplified form the expressions for the $`\omega 5\pi `$ amplitudes are not easy to use for evaluation of the branching ratios. To go further, it should be noted the following. One can check that the invariant mass of the $`4\pi `$ system on which the contribution of the diagrams shown in Fig. 1(a) depends, changes in very narrow range $`558\text{ MeV}<m_{4\pi }<642\text{ MeV}`$. Hence, one can set it in all the $`\rho `$ propagators standing as the first terms in all square brackets in Eqs. (21) and (18), with the accuracy $`20\%`$ in width, to the equilibrium value $`\overline{m_{4\pi }^2}^{1/2}=620`$ MeV evaluated for the pion energy $`E_\pi =m_\omega /5`$ which gives the dominant contribution. The same is true for the invariant mass of the pion pairs on which the $`\rho `$ propagators standing as the last terms in square brackets of the above expressions depend. This invariant mass varies in the narrow range $`280\text{ MeV}<m_{2\pi }<360\text{ MeV}`$. With the same accuracy, one can set it to $`\overline{m_{2\pi }^2}^{1/2}=295`$ MeV in all relevant propagators. On the other hand, the amplitude of the process $`\omega \rho ^0\pi ^0(2\pi ^+2\pi ^{})\pi ^0`$ is
$`M[\omega \rho ^0\pi ^0(2\pi ^+2\pi ^{})\pi ^0]`$ $`=`$ $`4{\displaystyle \frac{N_cg_{\rho \pi \pi }g^2}{8(2\pi )^2f_\pi ^3}}`$ (26)
$`\times \epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu (q_1+q_2)_\lambda `$
$`\times {\displaystyle \frac{q_{4\sigma }}{D_\rho (qq_4)}},`$
where the momentum assignment is the same as in Eq. (18). The other relevant amplitude corresponding to the first diagram in Fig. 1(b) is
$`M[\omega \rho ^0\pi ^0(\pi ^+\pi ^{})(\pi ^+\pi ^{}\pi ^0)]`$ $`=`$ $`{\displaystyle \frac{N_cg_{\rho \pi \pi }g^2}{8(2\pi )^2f_\pi ^3}}\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu `$ (28)
$`\times (1+P_{12})(1+P_{35}){\displaystyle \frac{q_{1\lambda }q_{3\sigma }}{D_\rho (q_1+q_3)}}`$
Then, taking into account the above consideration concerning the invariant masses, and comparing Eqs. (18), (26), and (28), one can see that
$`M(\omega 2\pi ^+2\pi ^{}\pi ^0)`$ $``$ $`{\displaystyle \frac{5}{2}}M[\omega \rho ^0\pi ^0(2\pi ^+2\pi ^{})\pi ^0]`$ (30)
$`\times \left[1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right].`$
The same treatment shows that
$`M(\omega \pi ^+\pi ^{}3\pi ^0)`$ $``$ $`{\displaystyle \frac{5}{2}}M[\omega \rho ^+\pi ^{}(\pi ^+3\pi ^0)\pi ^{}]`$ (32)
$`\times \left[1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right],`$
where
$`M[\omega \rho ^+\pi ^{}(\pi ^+3\pi ^0)\pi ^{}]`$ $`=`$ $`4{\displaystyle \frac{N_cg_{\rho \pi \pi }g^2}{8(2\pi )^2f_\pi ^3}}`$ (34)
$`\times {\displaystyle \frac{\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu q_{1\lambda }q_{2\sigma }}{D_\rho (qq_2)}},`$
and the final momenta assignment is the same as in Eq. (21). The numerical values of $`\overline{m_{4\pi }^2}^{1/2}`$ and $`\overline{m_{2\pi }^2}^{1/2}`$ found above are such that the correction factor in parentheses of Eqs. (28) and (30) amounts to $`20\%`$ in magnitude. In what follows, the above correction will be taken into account as an overall factor of 0.64 in front of the branching ratios of the decays $`\omega 5\pi `$. When making this estimate, the imaginary part of the $`\rho `$ propagators in square brackets of Eq. (28) and (30) is neglected. This assumption is valid with the accuracy better than $`1\%`$ in width.
It would be useful to fulfill the model estimate of the $`\omega 5\pi `$ partial widths as follows. The corresponding equilibrium pion momenta are $`|๐ช_{\pi ^+}|=70`$ MeV and $`|๐ช_{\pi ^0}|=79`$ MeV. The integrations over angles of final pions can be fulfilled assuming them independent. Using the nonrelativistic expression for phase space of five pions ,
$$R_5=\frac{\pi ^6(\underset{i}{\overset{5}{}}m_{\pi i})^{1/2}}{60(_i^5m_{\pi i})^{3/2}}(m_\omega \underset{i}{\overset{5}{}}m_{\pi i})^5$$
(35)
whose numerical value coincides with the accuracy $`1\%`$ with the numerically evaluated exact expression, and introducing the branching ratio at the $`\omega `$ mass as
$$B_{\omega 5\pi }=\mathrm{\Gamma }_{\omega 5\pi }(m_\omega )/\mathrm{\Gamma }_\omega ,$$
(36)
one finds
$`B(\omega 5\pi )`$ $``$ $`\left[{\displaystyle \frac{5N_c}{2\pi ^2}}\left({\displaystyle \frac{g_{\rho \pi \pi }g^2}{8f_\pi ^3}}\right){\displaystyle \frac{m_\omega |๐ช_{\pi ^+}|}{|D_\rho (\overline{m_{4\pi }^2})|}}\right]^2`$ (39)
$`\times {\displaystyle \frac{R_5}{18(2\pi )^{11}m_\omega \mathrm{\Gamma }_\omega }}\left|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right|^2`$
$`\times \{{\displaystyle \genfrac{}{}{0pt}{}{|๐ช_{\pi ^0}|^2(2\pi ^+2\pi ^{}\pi ^0)}{\frac{|๐ช_{\pi ^+}|^2}{3}(\pi ^+\pi ^{}3\pi ^0)}}`$
The calculation gives $`B(\omega 2\pi ^+2\pi ^{}\pi ^0)=2.5\times 10^9`$ and $`B(\omega \pi ^+\pi ^{}3\pi ^0)=1.0\times 10^9`$.
The evaluation of the partial widths valid with accuracy $`20\%`$ can be obtained upon taking the amplitude of each considered decays as 5/2 times the $`\rho \pi `$ production state amplitude with the subsequent decay $`\rho 4\pi `$, and calculate the partial width using the calculated widths of the latter:
$`B_{\omega 2\pi ^+2\pi ^{}\pi ^0}`$ $`=`$ $`\left|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right|^2\left({\displaystyle \frac{5}{2}}\right)^2{\displaystyle \frac{2}{\pi \mathrm{\Gamma }_\omega }}`$ (43)
$`\times {\displaystyle _{4m_{\pi ^+}}^{m_\omega m_{\pi ^0}}}dm`$
$`\times {\displaystyle \frac{m^2\mathrm{\Gamma }_{\omega \rho ^0\pi ^0}(m)\mathrm{\Gamma }_{\rho 2\pi ^+2\pi ^{}}(m)}{|D_\rho (m^2)|^2}}`$
$`=1.1\times 10^9`$
where
$$\mathrm{\Gamma }_{\omega \rho ^0\pi ^0}(m)=g_{\omega \rho \pi }^2q^3(m_\omega ,m,m_{\pi ^0})/12\pi ,$$
$$g_{\omega \rho \pi }=\frac{N_cg^2}{8\pi ^2f_\pi }=14.3\text{ GeV}^1.$$
Note also the $`a^1`$ dependence of the $`\omega 5\pi `$ width on the HLS parameter $`a`$. The branching ratio $`B_{\omega \pi ^+\pi ^{}3\pi ^0}`$ is obtained from Eq. (43) upon inserting the lower integration limit to $`m_{\pi ^+}+3m_{\pi ^0}`$, $`m_{\pi ^0}m_{\pi ^+}`$ in the expression for the momentum $`q`$ and substitution of the $`\rho ^+\pi ^+3\pi ^0`$ decay width corrected for the mass difference of charged and neutral pions. Of course, the main correction of this sort comes from the phase space volume of the final 4$`\pi `$ state. One obtains
$`B_{\omega \pi ^+\pi ^{}3\pi ^0}`$ $`=`$ $`\left|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right|^2\left({\displaystyle \frac{5}{2}}\right)^2{\displaystyle \frac{2}{\pi \mathrm{\Gamma }_\omega }}`$ (47)
$`\times {\displaystyle _{m_{\pi ^+}+3m_{\pi ^0}}^{m_\omega m_{\pi ^+}}}dm`$
$`\times {\displaystyle \frac{m^2\mathrm{\Gamma }_{\omega \rho ^+\pi ^{}}(m)\mathrm{\Gamma }_{\rho ^+\pi ^+3\pi ^0}(m)}{|D_\rho (m^2)|^2}}`$
$`=8.5\times 10^{10}`$
where
$$\mathrm{\Gamma }_{\omega \rho ^+\pi ^{}}(m)=g_{\omega \rho \pi }^2q^3(m_\omega ,m,m_{\pi ^+})/12\pi .$$
As is pointed out in Ref. , the inclusion of the direct $`\omega \pi ^+\pi ^{}\pi ^0`$ vertex reduces the 3$`\pi `$ decay width of the $`\omega `$ by $`33\%`$. This implies that one should make the following replacement to take into account the effect of the pointlike diagrams in Fig. 1(b) in the expression for the suppression factor:
$`\left|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2D_\rho (\overline{m_{2\pi }^2})}}\right|^2`$ $``$ $`|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{2}}[{\displaystyle \frac{1}{D_\rho (\overline{m_{2\pi }^2})}}`$ (50)
$`{\displaystyle \frac{1}{3m_\rho ^2}}\left]\right|^2|1{\displaystyle \frac{D_\rho (\overline{m_{4\pi }^2})}{3D_\rho (\overline{m_{2\pi }^2})}}|^2`$
$`0.75,`$
instead of 0.64, which results in the increase of the above branching ratios by the factor of 1.17.
The numerical value of the $`\omega 5\pi `$ decay width changes by the factor of two when varying the energy within $`\pm \mathrm{\Gamma }_\omega /2`$ around the $`\omega `$ mass. In other words, the dependence of this partial width on energy is very strong. This is illustrated by Fig. 2 where the $`\omega 5\pi `$ excitation curves in $`e^+e^{}`$ annihilation,
$`\sigma _{e^+e^{}\omega 5\pi }(s)`$ $`=`$ $`12\pi \left({\displaystyle \frac{m_\omega }{E}}\right)^3\mathrm{\Gamma }_{\omega e^+e^{}}(m_\omega )`$ (52)
$`\times {\displaystyle \frac{\mathrm{\Gamma }_\omega B_{\omega 5\pi }(E)}{\left[(sm_\omega ^2)^2+(m_\omega \mathrm{\Gamma }_\omega )^2\right]}},`$
are plotted. Here $`B_{\omega 2\pi ^+2\pi ^{}\pi ^0}(E)`$ \[$`B_{\omega \pi ^+\pi ^{}3\pi ^0}(E)`$\] is given by Eq. (43) \[(47)\], respectively, with the substitution $`m_\omega E`$. The mentioned strong energy dependence of the partial width results in the asymmetric shape of the $`\omega `$ resonance and the shift of its peak by $`+0.7`$ MeV . As is seen from Fig. 2, the peak value of the $`5\pi `$ state production cross section is about 1.5-2.0 femtobarns. Yet the decays $`\omega 5\pi `$ can be observable on $`e^+e^{}`$ colliders. Indeed, with the luminosity $`L=10^{33}\text{cm}^2\text{s}^1`$ near the $`\omega `$ peak, which seems to be feasible, one may expect about 2 events per week for the considered decays to be detected at these colliders.
As for the angular distributions of the final pions are concerned, they, of course, should be deduced from the full amplitudes Eqs. (18) and (21). However, some qualitative conclusions about the angular distributions can be drawn from the simplified expressions Eqs. (26), (30). Since helicity is conserved, only the states of the $`\omega (782)`$ with the spin projections $`\lambda =\pm 1`$ on the $`e^+e^{}`$ beam axes given by the unit vector $`๐ง_0`$ are populated. In what follows the suitable notation for the vector product of the pion momenta are used:
$`[๐ช_i\times ๐ช_j]`$ $`=`$ $`|๐ช_i||๐ช_j|\mathrm{sin}\theta _{ij}`$ (54)
$`\times (\mathrm{sin}\mathrm{\Theta }_{ij}\mathrm{cos}\mathrm{\Phi }_{ij},\mathrm{sin}\mathrm{\Theta }_{ij}\mathrm{sin}\mathrm{\Phi }_{ij},\mathrm{cos}\mathrm{\Theta }_{ij}).`$
In other words, $`\theta _{ij}`$ is the angle between the pion momenta $`๐ช_i`$ and $`๐ช_j`$, $`\mathrm{\Theta }_{ij}`$, $`\mathrm{\Phi }_{ij}`$ being the polar and azimuthal angles of the normal to the plane spanned by the momenta $`๐ช_i`$ and $`๐ช_j`$. Choosing $`๐ง_0`$ to be the unit vector along z axes, the probability density of the emission of two $`\pi ^+`$โs with the momenta $`๐ช_1`$, $`๐ช_2`$, and $`\pi ^0`$ with the momentum $`๐ช_4`$ is represented as
$`w`$ $``$ $`\left[๐ช_4\times \left(๐ช_1+๐ช_2\right)\right]^2\left(๐ง_0\left[๐ช_4\times \left(๐ช_1+๐ช_2\right)\right]\right)^2`$ (58)
$`=๐ช_4^2[๐ช_1^2\mathrm{sin}^2\theta _{41}\mathrm{sin}^2\mathrm{\Theta }_{41}+๐ช_2^2\mathrm{sin}^2\theta _{42}\mathrm{sin}^2\mathrm{\Theta }_{42}`$
$`+2|๐ช_1||๐ช_2|\mathrm{sin}\mathrm{\Theta }_{41}\mathrm{sin}\mathrm{\Theta }_{42}\mathrm{sin}\theta _{41}\mathrm{sin}\theta _{42}`$
$`\times \mathrm{cos}(\mathrm{\Phi }_{41}\mathrm{\Phi }_{42})]`$
in the case of the final state $`2\pi ^+2\pi ^{}\pi ^0`$. Here the momentum assignment is the same as in Eq. (18). The angular distribution of two $`\pi ^{}`$โs with the momenta $`๐ช_3`$, $`๐ช_5`$, and $`\pi ^0`$ is obtained from Eq. (58) upon the replacement $`๐ช_{1,2}๐ช_{3,5}`$, because the identity $`\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu (q_1+q_2)_\lambda q_{4\sigma }=\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu (q_3+q_5)_\lambda q_{4\sigma }`$ is valid. Since another identity $`\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu (q_1+q_2)_\lambda q_{4\sigma }=\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu (q_1+q_2)_\lambda (q_3+q_5)_\sigma `$ is valid, one can write the angular distribution that includes four charged pions:
$`w`$ $``$ $`[(๐ช_1+๐ช_2)\times (๐ช_3+๐ช_5)]^2\left(๐ง_0[(๐ช_1+๐ช_2)\times (๐ช_3+๐ช_5)]\right)^2`$ (63)
$`=(1+P_{12})(1+P_{35})๐ช_1^2๐ช_3^2\mathrm{sin}^2\theta _{13}\mathrm{sin}^2\mathrm{\Theta }_{13}`$
$`+2|๐ช_1||๐ช_2|(1+P_{35})๐ช_3^2\mathrm{sin}\theta _{13}\mathrm{sin}\theta _{23}\mathrm{sin}\mathrm{\Theta }_{13}\mathrm{sin}\mathrm{\Theta }_{23}\mathrm{cos}(\mathrm{\Phi }_{13}\mathrm{\Phi }_{23})`$
$`+2|๐ช_3||๐ช_5|(1+P_{12})๐ช_1^2\mathrm{sin}\theta _{13}\mathrm{sin}\theta _{15}\mathrm{sin}\mathrm{\Theta }_{13}\mathrm{sin}\mathrm{\Theta }_{15}\mathrm{cos}(\mathrm{\Phi }_{13}\mathrm{\Phi }_{15})`$
$`+2|๐ช_1||๐ช_2||๐ช_3||๐ช_5|(1+P_{35})\mathrm{sin}\theta _{13}\mathrm{sin}\theta _{25}\mathrm{sin}\mathrm{\Theta }_{13}\mathrm{sin}\mathrm{\Theta }_{25}\mathrm{cos}(\mathrm{\Phi }_{13}\mathrm{\Phi }_{25}).`$
Here $`P_{ij}`$ interchanges the indices $`i`$ and $`j`$. In the case of the final state $`\pi ^+\pi ^{}3\pi ^0`$ the corresponding probability density can be obtained from Eqs. (32) and (34) and looks as
$`w`$ $``$ $`[๐ช_1\times ๐ช_2]^2(๐ง_0[๐ช_1\times ๐ช_2])^2`$ (65)
$`=๐ช_1^2๐ช_2^2\mathrm{sin}^2\theta _{21}\mathrm{sin}^2\mathrm{\Theta }_{21}.`$
Here the momentum assignment is the same as in Eq. (21). The corresponding angular distribution of one charged, say $`\pi ^+`$, and three neutral pions can be obtained from Eqs. (32) and (34) upon using the identity $`\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu q_{1\lambda }q_{2\sigma }=\epsilon _{\mu \nu \lambda \sigma }q_\mu ฯต_\nu q_{1\lambda }(q_3+q_4+q_5)_\sigma `$ and looks as
$`w`$ $``$ $`\left[๐ช_1\times {\displaystyle \underset{i}{}}๐ช_i\right]^2\left(๐ง_0\left[๐ช_1\times {\displaystyle \underset{i}{}}๐ช_i\right]\right)^2`$ (69)
$`=๐ช_1^2[{\displaystyle \underset{i}{}}๐ช_i^2\mathrm{sin}^2\theta _{i1}\mathrm{sin}^2\mathrm{\Theta }_{i1}`$
$`+2{\displaystyle \underset{ij}{}}|๐ช_i||๐ช_j|\mathrm{sin}\theta _{i1}\mathrm{sin}\theta _{j1}\mathrm{sin}\mathrm{\Theta }_{i1}\mathrm{sin}\mathrm{\Theta }_{j1}`$
$`\times \mathrm{cos}(\mathrm{\Phi }_{i1}\mathrm{\Phi }_{j1})].`$
Here indeces $`i,j`$ run over 3,4,5.
The strong energy dependence of the five pion partial width of the $`\omega `$ implies that the branching ratio at the $`\omega `$ mass, Eq. (36), evaluated above, is slightly different from that determined by the expression
$$B_{\omega 5\pi }^{\mathrm{aver}}(E_1,E_2)=\frac{2}{\pi }_{E_1}^{E_2}๐E\frac{E^2\mathrm{\Gamma }_\omega B_{\omega 5\pi }(E)}{(E^2m_\omega ^2)^2+(m_\omega \mathrm{\Gamma }_\omega )^2}.$$
(70)
Taking $`E_1=`$ 772 MeV and $`E_2=`$ 792 MeV, one finds $`B_{\omega 2\pi ^+2\pi ^{}\pi ^0}^{\mathrm{aver}}(E_1,E_2)=9.0\times 10^{10}`$ and $`B_{\omega \pi ^+\pi ^{}3\pi ^0}^{\mathrm{aver}}(E_1,E_2)=6.7\times 10^{10}`$ to be compared to Eq. (43) and (47), respectively. In particular, the quantity $`B_{\omega 2\pi ^+2\pi ^{}\pi ^0}^{\mathrm{aver}}(E_1,E_2)`$ is the relevant characteristics of this specific decay mode in photoproduction experiments. The Jefferson Lab โphoton factoryโ could also be suitable for detecting the five pion decays of the $`\omega `$. However, in view of the suppression of the $`\omega `$ photoproduction cross section by the factor of 1/9 as compared with the $`\rho `$ one, the total number of $`\omega `$ mesons will amount to $`7\times 10^8`$ per nucleon. Hence, the increase of intensity of this machine by the factor of 50 is highly desirable, in order to observe the decay $`\omega 5\pi `$ and measure its branching ratio. Evidently, the $`\omega `$ photoproduction on heavy nuclei is preferable in view of the dependence of the cross section on atomic weight $`A`$ growing as $`A^{0.80.95}`$ .
The conclusions about the angular distributions in photoproduction are the following. Of course, their general expression should be deduced from the full decay amplitudes Eqs. (18) and (21), together with the detailed form of the photoproduction mechanism. The qualitative picture, however, can be obtained upon noting that $`s`$-channel helicity conservation is a good selection rule for the photoproduction reactions. Then in the helicity reference frame characterized as the frame where the $`\omega `$ is at rest, while its spin quantization axes is directed along the $`\omega `$ momentum in the center-of-mass system, the expressions for the angular distributions coincide with the expressions Eqs. (58),(63), (65) and (69). Since, at high energies, the direction of the final $`\omega `$ momentum lies at the scattering angle less than $`0.5^{}`$ in the case of the photoproduction on heavy nuclei, the vector $`๐ง_0`$ can be treated as pointed along the photon beam direction.
Together with the $`e^+e^{}`$ annihilation experiments, the study of the photoproduction of the five pion states on heavy nuclei would also allow to measure the corresponding partial width of the $`\omega (782)`$. The comparison with theoretical expectations presented here would give the possibility of testing the predictions of chiral models that include the vector mesons in the situation where the decay amplitudes are determined by very low pion momenta.
We are grateful to G. N. Shestakov for discussion. The present work is supported in part by the grant RFBR-INTAS IR-97-232. |
warning/0003/cond-mat0003173.html | ar5iv | text | # Distribution function of the local density of states of a oneโchannel weakly disordered ring in an external magnetic field
## I Introduction
Interference effects in lowโdimensional disordered conductors still attract attention from both experimental and theoretical physicists, although all main features and a lot of new effects have already been discovered during the last twenty years. The prediction of the oscillation in the kinetic coefficients in multiply connected disordered normal metals in an external magnetic field and its experimental observation in a Mg cylinder was a very excellent examination of weak localization phenomena, since the coherence of the electron wavefunction during the circulation of a closed contour is required to observe an oscillation. The period of oscillation of the magnetoresistance predicted and observed firstly was equal to half of the magnetic flux quantum $`\varphi _0=\frac{hc}{e}`$. Further improvements of the experiments on rings with large diameters and small widths gave rise to the observation of a magnetoresistance oscillation with the period $`\varphi _0`$\[\], furthermore in the experiment of Chandrasekhar et al. both periods were observed. Such a complex magnetic field dependence of the magnetoresistance seems to be related with the statistical properties of the sample.
Another development in theory was the prediction of a persistent current in a oneโchannel disordered isolated loop. Due to the analogy between loop and oneโdimensional (1D) lattice with a period equal to the circumference of the loop, a circulating current in rings was suggested which is a periodic function of the enclosed flux with a period of $`\varphi _0`$. Studies of the effects on the persistent current at finite but small temperature and weak inelastic scattering show that both weak inelastic and elastic scattering do not destroy it. In experiments, the persistent current was also observed. A magnetization measurement was performed in \[\] on $`N=10^7`$ disconnected copper loops at $`T<1.5\mathrm{K}`$ where the electron phase coherence length $`L_\phi `$ exceeds the loopโs circumference $`L=2.2\mu \mathrm{m}`$, and it shows evidence for a fluxโperiodic persistent current with halved period and $`3\times 10^3\frac{ev_F}{L}`$ amplitude per ring, which is remarkably higher than the theoretically expected value for the persistent current per ring, $`\sqrt{\frac{l}{LN}}\frac{ev_F}{L}`$ \[\]. Measurements of the persistent current on single loops (with at least a few channels) in the diffusive ($`L=8\mu \mathrm{m}`$ and $`l=70\mathrm{n}\mathrm{m}`$, where $`l`$ is the elastic mean free path) and ballistic ($`L=8.5\mu \mathrm{m}`$ and $`l=11\mu \mathrm{m}`$) regimes reveal the period $`\varphi _0`$. The amplitude of the harmonic with $`\frac{\varphi _0}{2}`$ was measured to be smaller by a factor of 2-3 than that of the $`\varphi _0`$ harmonic in \[\].
Effects of impurities in most of the theoretical investigations were taken into account by the transfer matrix method according to the Landauer expression and by generalizations of this method to $`n`$โchannel systems. Although the Landauer formula gives fullโflux periodicity for all physical parameters, averaging over ensembles of rings or the calculation of the dynamical current instead of the thermodynamical potential were suggested as an explanation of the observed halved periodicity. In the process of averaging over different impurity realizations in the ensemble, the number of particles in each ring is proposed to be constant, i.e. the persistent current is assumed to be determined by the thermodynamic potential instead of the grand canonical potential. Although there has been done a lot of work on the AharonovโBohm effect, the existing theories of nonโinteracting electrons still can explain neither the high value of the experimentally observed persistent currents nor its diamagnetic sign. This can be partially connected with the complexity of the experiments, particularly with difficulties of the separation of phase effects in the rings with relatively large width (larger than the mean free path) from orbital ones.
On the other hand, correlation effects may be a reason for the discrepancy between theory and experiment. Unfortunately, there is still no agreement on the effects of Coulomb interaction on the amplitude of the persistent current. Studies based on spinless electrons in 1D continuum and lattice models gave controversial results, so the amplitude of the persistent current was shown to be increased up to its disorderโfree value according to the former model, but a MottโHubbard metalโinsulator transition in the latter model was found to reduce the amplitude, or at least a possible increase of the amplitude was negligibly small. There was also the suggestion that correlation can change the fundamental period with the magnetic flux and create fractional periodicity in a 1D ring. Impurities and correlation acting together are again subject of controversy. Thus, considering weak localization corrections in first order in the electronโelectron interaction to the grand canonical potential, a persistent current with a period of $`\frac{\varphi _0}{2}`$ and an amplitude of $`\frac{ev_Fl}{L^2}`$, corresponding to the experiment, was obtained, while Monte Carlo simulation on a 1D Luttinger liquid resulted in a persistent current with period $`\varphi _0`$ and with an amplitude decreased through interaction. Resumming the existing results it can be said that neither the noninteracting electron model nor models of correlated electrons yet gave satisfactory answers to the questions put forward by the experiments. These concern the period (under what condition both periods or only the halved period are observed), the amplitude of the persistent current and its diamagnetic sign (for the correlated electron model the diamagnetic sign requires an attractive interaction between the electrons).
To prevent the interference of the orbital effects in the presence of an external magnetic field with the phase effects, the width of the ring should be chosen as narrow as possible, i.e. a oneโchannel ring with random impurities seems to be an ideal tool to study the AharonovโBohm effect. However, interference effects in 1D disordered systems as a result of the coherent backscattering processes are strong irrespective of the degree of randomness. The diffusion approximation, which was used in previous studies, is not acceptable in 1D systems even for the case of weak disorder.
Approaching the problem thoroughly, we use in this paper a weakly disordered noninteracting electron model and construct for it a new exactly solvable diagrammatic method, which is an extension of Berezinskiiโs method to the problem with periodic boundary conditions. Within this model, we sum up all impurity scattering diagrams in the framework of the Born approximation.
In Sec.II, we describe the method. We calculate the density of electronic states (DOS) in Sec.III. Indeed an average value is not enough to describe the observable parameters in low dimensional mesoscopic systems. It is well known that the physical parameters of a mesoscopic system with size $`L`$ satisfying the condition $`l<LL_\phi `$ fluctuate from sample to sample, i.e. selfโaveraging is violated. At $`T=0`$ all sufficiently large systems become mesoscopic. In this case, high moments give a considerable contribution, which results in strong differences between average value and typical one of the observable parameter, i.e. the average value loses its significance to characterize the experimental observation. In Sec.IV the diagrammatical method is applied to find the $`k`$th moments of the DOS, $`\rho ^k(ฯต,\varphi )`$. The obtained equations for $`\rho ^k(ฯต,\varphi )`$ show that, in contrast to the average value of the DOS, all higher moments oscillate with the halved period, $`\frac{\varphi _0}{2}`$. Although the structure of the equations is complicated, the latter can be solved for the weak localization regime when the condition $`lL`$ is satisfied. This procedure is given in Sec.V.
The zeroth (not depending on $`\varphi `$) and the first (oscillating with $`\frac{\varphi _0}{2}`$) harmonic of $`\rho ^k(ฯต,\varphi )`$ are studied explicitely in this section. Both are shown to increase with $`k`$ as $`\mathrm{exp}(k^2)`$, which gives rise to a logarithmic normal distribution.
In section VI we conclude our results and discuss possibilities to extend our approach to related problems.
## II Description of the method
We consider here a oneโchannel disordered ring with length $`L=2\pi r`$, threaded by a magnetic flux $`\varphi `$ through the opening. The Hamiltonian of the system can be written as
$$H=\frac{\mathrm{}^2}{2m^{}r^2}\left(i\frac{}{\phi }+\frac{\varphi }{\varphi _0}\right)^2+V_{\mathrm{imp}}(\phi )$$
(1)
where $`\varphi _0=\frac{hc}{e}`$ is the fundamental period of a flux quantum and $`m^{}`$ is the effective mass of an electron. The potential $`V_{\mathrm{imp}}`$ of randomly distributed impurities is considered to be weak, so that scattering processes can be studied in the framework of the Born approximation. Below, we use the spatial variable $`x=r\phi `$ instead of the angle $`\phi `$.
Our aim in this section is to construct a diagrammatical method for the calculation of the average values of the DOS, $`\rho (ฯต,\varphi ;L)`$, and of its moments $`\rho ^n(ฯต,\varphi ;L)`$. The bracket $`\mathrm{}`$ denotes the average over the impurity configurations. Expressing the DOS by means of retarded ($`G_R`$) and advanced ($`G_A`$) Greenโs functions (GF) as
$$\rho (ฯต,\varphi ;x)=\frac{1}{\pi }\mathrm{Im}G_R(ฯต,\varphi ;x,x)=\frac{1}{2\pi i}[G_A(ฯต,\varphi ;x,x)G_R(ฯต,\varphi ;x,x)]$$
(2)
the $`n`$โth moment of $`\rho (ฯต,\varphi ;x)`$ can be given by
$$\rho ^k(ฯต,\varphi ;x)=\frac{1}{(2\pi i)^n}\underset{l=0}{\overset{k}{}}\left(\genfrac{}{}{0pt}{}{n}{k}\right)(1)^lG_R^l(ฯต,\varphi ;x,x)G_A^{kl}(ฯต,\varphi ;x,x)$$
(3)
The Berezinskii diagram technique is applied to calculate the average value of a single GF and higherโorder correlators $`G_R^lG_A^{kl}`$. In contrast to strictly 1D disordered wires, quantum corrections to the DOS of a ring turn out to exist even for the weakly disordered limit due to periodicity.
As for infinite systems, we consider as starting point a free particle with wave function $`\psi _p(x)\mathrm{exp}(ipx)`$, where the momentum can assume arbitrary values, leading to a continuous spectrum $`ฯต_p=\frac{\mathrm{}^2}{2m^{}}(p\frac{\varphi }{\varphi _0r})^2`$. The โbareโ GF $`G_{R,A}^0`$ can be calculated easily:
$$G_{R,A}^0(ฯต,\varphi ;x,x^{})=\frac{dp}{2\pi }\frac{e^{ip(xx^{})}}{ฯตฯต_p\pm i\eta }=\frac{i}{v(ฯต)\mathrm{}}e^{i2\pi \frac{\varphi }{\varphi _0}\frac{xx^{}}{L}\pm ip(ฯต)|xx^{}|\frac{\eta }{v(ฯต)}|xx^{}|}$$
(4)
where $`L`$ is the circumference of the ring and the parameter $`\eta `$ is introduced phenomenologically to model inelastic processes, which result in a blurring of the energy levels. $`v(ฯต)=\sqrt{\frac{2ฯต}{m^{}}}`$ and $`\mathrm{}p(ฯต)=\sqrt{2m^{}ฯต}`$ are the velocity and the momentum of an electron with energy $`ฯต`$, respectively. Notice that Zeeman splitting has not been taken into account here. For an electron with spin $`s=\pm \frac{1}{2}`$, it would throughout the paper lead to a shift of the energy as $`ฯตฯตsg\mu _BB`$ (where $`g`$ is the gyromagnetic ratio of the electron and $`\mu _B`$ is the Bohr magneton).
This bare GF, however, is not yet the real GF for an electron in a ring without impurities, since it does not reflect the finite size of the system and the periodic boundary conditions. These are taken into account by allowing the particles to make arbitrary revolutions around the ring, which leads to the expected quantization effect. According to this prescription, the GF for a clean ring $`\stackrel{~}{G}_{R,A}^0`$ is
$$\stackrel{~}{G}_{R,A}^0(ฯต,\varphi ;x,x^{})=G_{R,A}^0(ฯต,\varphi ;x,x^{})+G_{R,A}^0(ฯต,\varphi ;x,x^{}+L)+G_{R,A}^0(ฯต,\varphi ;x,x^{}L)+G_{R,A}^0(ฯต,\varphi ;x,x^{}+2L)\mathrm{}$$
(5)
One may verify this approach by calculating the DOS of a clean ring in a magnetic field $`\rho _0(ฯต,\varphi )`$ from Eqs.(2), (4), and (5):
$$\rho _0(ฯต,\varphi )=\rho _0+2\rho _0\underset{n=1}{\overset{\mathrm{}}{}}\mathrm{cos}(p(ฯต)Ln)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0}n)e^{\frac{\eta }{v(ฯต)}Ln}$$
(6)
where $`\rho _0=1/(\pi \mathrm{}v(ฯต))`$ is the DOS for a pure and infinite 1D system. As $`\eta 0`$, Eq.(6) displays the discrete behavior for the DOS of a clean ring.
In the diagrammatical technique, the retarded (advanced) GF of an electron moving in the field of randomly distributed impurities is represented by an ordinary (double) continuous line in real space, which goes from point $`x`$ to $`x^{}`$ after multiple scattering on a given impurity configuration, realized by the potentials $`V(x_i)`$ with impurities placed at the points $`\{x_i\}`$. For the DOS and its moments it suffices to consider $`x=x^{}`$. Since the โbareโ GFs \[Eq.(4)\] between two successive scatterings have factorable structure, the coordinate dependence can be transfered from the lines to the vertices. Averaging over the random Gaussian potential leads to a pairing of the impurity vertices. Their strength is measured by the inverse forward (backward) scattering length,
$$\frac{1}{l^+(ฯต)}=\frac{2}{\mathrm{}^2v^2(ฯต)}_0^{\mathrm{}}U(x)๐x\text{and}\frac{1}{l^{}(ฯต)}=\frac{2}{\mathrm{}^2v^2(ฯต)}_0^{\mathrm{}}U(x)๐x\mathrm{cos}(p(ฯต)x)$$
(7)
For the Born approximation to be applicable, the correlator $`U(xx^{})=V(x)V(x^{})`$ should have a width much smaller than the mean distance between impurities $`\frac{1}{c}`$, and $`\frac{1}{c}ล^\pm `$. In the extreme case of a white noise potential $`U(xx^{})=cU_0^2_{n=\mathrm{}}^{\mathrm{}}\delta (xx^{}+nL)`$, the two scattering lengths become equal, $`l^{}=l^+=2l`$. Given in Fig.1 are the essential vertices selected according to the condition $`p_Fl1`$, with $`p_F`$ and $`l`$ being the Fermi momentum and the mean free path, respectively. Although the โbareโ GFs depend on the direction due to the magnetic field, the internal vertices in Fig.1 do not differ from those of Berezinskii. All dependence on the magnetic field is transfered from the lines to the external vertices, which are shown in Fig.2.
As an example, a simple diagram contributing to $`G_R(ฯต,\varphi ;x,x)`$ is drawn in Fig.3a. For convenience, we cut the diagram at point $`x`$ and straighten the lines, which results in Fig.3b. Each diagram for a single GF is then characterized by $`m`$ pairs of lines returning to $`x`$ and $`n`$ through-going lines. Since the bulk of each diagram (i.e. after the removal of the external vertices) does not depend on the direction, $`n`$ is the sum of right-going $`n^+`$ and left-going $`n^{}`$ lines. For correlators $`G_R^l(ฯต,\varphi ;x,x)G_A^{kl}(ฯต,\varphi ;x,x)`$, we have to distinguish the number $`m`$ of returning lines on the l.h.s. and the number $`m^{}`$ of returning lines on the r.h.s., and in addition to introduce $`\overline{m}`$, $`\overline{m}^{}`$, and $`\overline{n}`$ for the advanced GF. An example for this case is shown in Fig.3c.
In contrast to Berezinskiiโs technique for an infinite 1D system, the external vertices depend on the direction. Also, the diagrams carry a factor $`\mathrm{exp}[2\pi i\frac{\varphi }{\varphi _0}(n^++\overline{n^+}n^{}\overline{n^{}})]`$ from the through-going lines. The number of pairs on the two sides of the cutting line, which we denote by $`m`$ and $`m^{}`$ for the retarded GF and by $`\overline{m}`$ and $`\overline{m}^{}`$ for the advanced GF, may in general differ by $`\pm 1`$ (i.e. $`mm^{}=1,0,1`$ and $`\overline{m}\overline{m}^{}=1,0,1`$) as in Sec. IV. However, for $`G_R(ฯต,\varphi ;x,x^{})`$ in the regime of weak disorder as in Sec. III, we have always $`m=m^{}`$.
## III The density of states
The diagrams for the DOS do not exhibit the full complexity presented in the previous section, since they contain only retarded GFs. Consequently, we can omit vertices d),e), and f) of Fig.1 and set $`m=m^{}`$. Following Berezinskiiโs method , we denote the sum of all diagrams with $`m`$ pairs of returning lines and $`n=n^++n^{}`$ through-going lines by $`Q_0(m,n;xx^{}=L)`$, where $`n^+`$($`n^{}`$) is the number of right-(left-)going lines that cross the whole diagram. Such a diagram is shown in Fig.3a and 3b. Since the magnetic field dependence has been extracted from $`Q_0`$, it only depends on the total number of through-going lines. The condition $`m=m^{}`$ restricts the possibilities to attach the external vertices to the cases a) and c) of Fig.4. The average value of the retarded GF can be expressed in terms of the kernel $`Q_0`$ as
$$\begin{array}{cc}\hfill G^+(ฯต,\varphi ;x,x)=& \frac{i}{v(ฯต)}\underset{m=0}{\overset{\mathrm{}}{}}\underset{n^+=0}{\overset{\mathrm{}}{}}\underset{n^{}=0}{\overset{\mathrm{}}{}}\left[\left(\genfrac{}{}{0pt}{}{m+n^+}{m}\right)\left(\genfrac{}{}{0pt}{}{m1+n^{}}{m1}\right)+\left(\genfrac{}{}{0pt}{}{m+n^{}}{m}\right)\left(\genfrac{}{}{0pt}{}{m1+n^+}{m1}\right)\delta _{m,0}\delta _{n^+,0}\delta _{n^{},0}\right]\hfill \\ & \mathrm{exp}\left\{ip(ฯต)L(n^++n^{})\frac{\eta L}{v(ฯต)}(n^++n^{})2\pi i\frac{\varphi }{\varphi _0}(n^+n^{})\right\}Q_0(m,n^++n^{};L)\hfill \end{array}$$
(8)
The two products of binomials in the bracket of Eq.(8) characterize the different possibilities to insert the $`n^+`$ right-going and the $`n^{}`$ left-going lines between $`m`$ pairs and correspond to the cases a) and c) of Fig.4, respectively. For $`m=n^+=n^{}=0`$, these two possibilities are degenerated into a pointโlike diagram. To avoid doubleโcounting in this case, the third term in the brackets has been added. The combinatorial factor, corresponding e.g. to the configuration in Fig.4a, can be obtained as follows: $`n^+`$ lines can be distributed at $`m+1`$ positions, before each of the loops on the l.h.s. and after the last loop, whereas the $`n^{}`$ leftโgoing lines can be inserted at $`m`$ positions before each of the loops on the r.h.s.. Denoting the number of lines at a given position with $`n_i^\pm `$, we have the restrictions $`n^+=n_1^++n_2^+\mathrm{}+n_{m+1}^+`$ and $`n^{}=n_1^{}+n_2^{}\mathrm{}+n_m^{}`$.
Summing over all these possibilities gives
$$\underset{\{n_i^+\}}{}\delta _{n^+,n_1^++n_2^++\mathrm{}+n_{m+1}^+}\underset{\{n_i^{}\}}{}\delta _{n^{},n_1^{}+n_2^{}+\mathrm{}+n_m^{}}=\left(\genfrac{}{}{0pt}{}{m+n^+}{m}\right)\left(\genfrac{}{}{0pt}{}{m1+n^{}}{m1}\right)$$
(9)
The expression for the DOS can be obtained by combining Eqs.(2) and (8):
$$\begin{array}{c}\hfill \rho (ฯต,\varphi )=\rho _0\underset{m=0}{\overset{\mathrm{}}{}}\underset{n=0}{\overset{\mathrm{}}{}}\underset{k=0}{\overset{n}{}}\left[2\left(\genfrac{}{}{0pt}{}{m+k}{m}\right)\left(\genfrac{}{}{0pt}{}{m+nk1}{m1}\right)\delta _{m,0}\delta _{n,0}\right]\\ \hfill \mathrm{cos}(p(ฯต)Ln)\mathrm{exp}(\frac{\eta }{v(ฯต)}Ln)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0}(2kn))Q_0(m,n;L)\end{array}$$
(10)
To find $`Q_0(m,n;xx^{})`$, we shift $`x`$ infinitesimally and examine the different impurity vertices that pass through the point $`x`$. The result is a differential equation for $`Q_0`$,
$$\frac{\mathrm{d}}{\mathrm{d}x}Q_0(m,n;x)=\left[\frac{(2m+n)^2}{2l^+}+\frac{n}{2l^{}}+\frac{m(m+n)}{l^{}}\right]Q_0(m,n;x)\frac{1}{l^{}}m^2Q_0(m1,n+2;x)$$
(11)
where the vertices a),b), and c) in Fig.1 contribute. For a) the number of possibilities to be included is $`2m+n`$, for b) $`\frac{1}{2}(2m+n)(2m+n1)`$ and to include c) there are $`m(m+n1)`$ ways that do not change $`m`$ and $`n`$ and $`m^2`$ possibilities that decrease $`m`$ by 1 pair and increase $`n`$ by 2. The process of construction of this equation is illustrated in Fig.5.
The kernel $`Q_0`$ satisfies the boundary condition
$$Q_0(m,n;xx^{}=L=0)=\delta _{m,0}$$
(12)
which expresses the absence of scattering for a ring with infinitesimal small circumference. The Eq.(11) for $`Q_0(m,n;x)`$ can be solved exactly as shown in Appendix A:
$$\begin{array}{cc}\hfill Q_0(m,n;L)=& \mathrm{exp}\left\{\frac{L}{2l^+}(2m+n)^2\frac{L}{l^{}}m(m+n)\frac{L}{2l^{}}n\right\}\hfill \\ & \underset{j=0}{\overset{m}{}}(1)^j\left(\genfrac{}{}{0pt}{}{m}{j}\right)\frac{m!(j+n2)!}{(m+j+n1)!}(2j+n1)\mathrm{exp}\left\{\frac{L}{l^{}}j(j+n1)\right\}\hfill \end{array}$$
(13)
Equations (10) and (13) give a complete description of the DOS for a oneโchannel ring in an external magnetic field. They are exact in the regime of weak disorder, when the condition $`p_Fl^\pm 1`$ or $`ฯต_F\tau 1`$ is satisfied ($`p_F`$ and $`ฯต_F`$ are Fermi momentum and Fermi energy, respectively). Further, since the circumference $`L`$ of the ring may vary within a large range ($`Ll^\pm `$ to $`l^\pm L<\mathrm{}`$), the results cover both the weak localization and the ballistic regimes. For the weak localization regime, when $`L\mathrm{max}(l^+,l^{})`$ the amplitude of $`Q_0(m,n;L)`$ decreases rapidly with $`m`$ and $`n`$ due to the exponential prefactor. Keeping harmonics up to $`n=2`$ in Eq.(10) and using Eq.(A6), we obtain the leading behavior for $`\rho (ฯต,\varphi )`$
$$\begin{array}{cc}\hfill \rho (ฯต,\varphi )=& \rho _0[1\frac{2L}{l^{}}\mathrm{exp}(\frac{2L}{l^+}\frac{L}{l^{}})]+2\rho _0\mathrm{exp}(\frac{L}{2l^+}\frac{L}{2l^{}})[\mathrm{cos}(p(ฯต)L)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0})e^{\frac{\eta L}{v(ฯต)}}\hfill \\ & +e^{\frac{3L}{2l^+}\frac{L}{2l^{}}}\mathrm{cos}(2p(ฯต)L)\mathrm{cos}(4\pi \frac{\varphi }{\varphi _0})e^{\frac{2\eta }{v(ฯต)}L}]\hfill \end{array}$$
(14)
We see that the main quantum correction to the average value of the DOS oscillates with a period of $`\varphi _0`$. The amplitude of this contribution decreases exponentially with the impurity strength or with increasing $`L`$, so that $`\rho =\rho _0`$ for $`L\mathrm{}`$.
The ballistic regime is realized for $`L\mathrm{min}(l^+,l^{})`$. Keeping terms up to first order in $`m`$, the DOS can be approximated in this limit by
$$\begin{array}{c}\hfill \rho (ฯต,\varphi )=\rho _0(ฯต,\varphi )\rho _0\frac{L}{l^+}\underset{n=0}{\overset{N^+}{}}n^2\mathrm{cos}(p(ฯต)Ln)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0}n)\rho _0\frac{L}{l^{}}\underset{n=0}{\overset{N^{}}{}}n\mathrm{cos}(p(ฯต)Ln)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0}n)\\ \hfill 2\rho _0\frac{L}{l^{}}\underset{n=0}{\overset{N^{}}{}}\underset{k=0}{\overset{n}{}}(k+1)\mathrm{cos}(p(ฯต)Ln)\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0}(2kn))\end{array}$$
(15)
where $`N^+\left[\sqrt{\frac{2l^+}{L}}\right]`$ and $`N^{}\left[\frac{2l^{}}{L}\right]`$, and the DOS of a clean ring $`\rho _0(ฯต,\varphi )`$ is given by Eq.(6). In the ballistic regime, $`L`$ is of the same order of magnitude as $`l^\pm `$, hence $`N^\pm `$ may be rather small integers. Therefore the oscillation with the full flux quantum $`\varphi _0`$ dominates in the ballistic regime.
In the absence of backward scattering ($`l^{}=\mathrm{}`$), $`Q_0(m,n;L)`$ is greatly simplified:
$$Q_0(m,n;L)=\mathrm{exp}\left\{\frac{L}{2l^+}n^2\right\}\delta _{m,0}$$
(16)
Substituting Eq.(16) in (10) and using $`lim_{m0}\left(\genfrac{}{}{0pt}{}{m+nk1}{m1}\right)=\delta _{n,k}`$, we can express the DOS as
$$\begin{array}{c}\hfill \rho (ฯต,\varphi )=\rho _0+\frac{\rho _0}{2}\sqrt{\frac{l^+}{2\pi L}}_{\mathrm{}}^{\mathrm{}}\mathrm{d}ze^{\frac{l^+}{2L}z^2}\{\frac{1}{\mathrm{exp}(ip(ฯต)Li2\pi \varphi /\varphi _0+iz+\frac{\eta }{v(ฯต)}L)1}\\ \hfill +\frac{1}{\mathrm{exp}(ip(ฯต)L+i2\pi \varphi /\varphi _0+iz+\frac{\eta }{v(ฯต)}L)1}+c.c\}\end{array}$$
(17)
From Eq.(17) one sees that forward scattering coherently shifts all energy levels. The value of this shifting is random with Gaussian distribution, with a typical value of $`\frac{\mathrm{}}{\tau ^+}\sqrt{\frac{l^+}{L}}`$, where $`\tau ^+`$ is the relaxation time due to forward scattering. The level repulsion in 1D disordered systems therefore is only due to backward scattering. The width of this level broadening of the averaged system is much smaller than Dingle broadening for the weak localization regime, whereas the two mechanisms can have comparable effects in the ballistic regime.
## IV Higher moments of the DOS and distribution functions
Exact calculations show that the average value of the DOS oscillates with a period of the flux quantum $`\varphi _0`$. To understand the reason for the experimentally observed oscillation of the persistent current in a sufficiently large ring ($`Ll`$) with the halved period, we calculate here higher moments of the distribution of the DOS. According to Eq.(3), we have to determine the correlators $`G_R^lG_A^{kl}`$ for the $`k`$th moment. In contrast to the Berezinskii technique for strictly 1D systems, the correlators here are characterized by only one block $`Q`$. An example is shown in Fig.3c. Each diagram contributing to $`G_R^lG_A^{kl}`$ consists of $`l`$ retarded and $`kl`$ advanced lines. For the $`i`$th retarded (advanced) GF, we count the number of left loops $`m_i`$($`\overline{m}_i`$), of right loops $`m_i^{}`$($`\overline{m}_i^{}`$), and of left- and right-going traversing lines $`n_i=n_i^++n_i^{}`$($`\overline{n}_i=\overline{n^+}_i+\overline{n^{}}_i`$). (Notice that the index $`i`$ of $`n_i`$ here denotes the number of the GF, whereas it was used for the position within one GF in the previous section.) For the different fermion lines, we now can attach the external vertices in four different ways, as shown in Fig.4. This allows $`m_im_i^{}=1,0,1`$ ($`\overline{m}_i\overline{m}_i^{}=1,0,1`$) for the each retarded (advanced) GF. It turns out that the block $`Q`$ depends only on the total numbers $`m=m_1+m_2+\mathrm{}+m_l`$, $`\overline{m}=\overline{m}_1+\overline{m}_2+\mathrm{}+\overline{m}_{kl}`$ and similarly $`m^{}`$,$`\overline{m}^{}`$,$`n`$,$`\overline{n}`$\[\]. It is clear from these considerations that the difference between $`m`$ and $`m^{}`$ is always restricted by $`lmm^{}l`$. This is automatically taken into account by the mixing coefficient $`\phi _l(m,m^{};n^+,n^{})`$. This coefficient is the generalization of the term in brackets of Eq.(8). It counts the different possibilities to attach external vertices (see Fig.4) and to distribute the through-going lines between the loops. Using Eq.(9), we can write it as
$$\begin{array}{cc}\hfill \phi _l(m,m^{};& n^+,n^{})=_{\{m_i\}}\delta _{m,m_1+m_2+\mathrm{}+m_l}_{\{m_i^{}\}}\delta _{m^{},m_1^{}+m_2^{}+\mathrm{}+m_l^{}}_{\{n_i^+\}}\delta _{n^+,n_1^++n_2^++\mathrm{}+n_l^+}_{\{n_i^{}\}}\delta _{n^{},n_1^{}+n_2^{}+\mathrm{}+n_l^{}}\hfill \\ & \underset{i=1}{\overset{l}{}}\{e^{2ip(ฯต)x\frac{2\eta }{v(ฯต)}x}\delta _{m_i,m_i^{}+1}\left(\genfrac{}{}{0pt}{}{m_i1+n_i^+}{m_i1}\right)\left(\genfrac{}{}{0pt}{}{m_i1+n_i^{}}{m_i1}\right)+e^{2ip(ฯต)x+\frac{2\eta }{v(ฯต)}x}\delta _{m_i,m_i^{}1}\left(\genfrac{}{}{0pt}{}{m_i+n_i^+}{m_i}\right)\left(\genfrac{}{}{0pt}{}{m_i+n_i^{}}{m_i}\right)\hfill \\ & +\delta _{m_i,m_i^{}}[\left(\genfrac{}{}{0pt}{}{m_i+n_i^+}{m_i}\right)\left(\genfrac{}{}{0pt}{}{m_i1+n_i^{}}{m_i1}\right)+\left(\genfrac{}{}{0pt}{}{m_i1+n_i^+}{m_i1}\right)\left(\genfrac{}{}{0pt}{}{m_i+n_i^{}}{m_i}\right)\delta _{m_i,0}\delta _{n_i^+,0}\delta _{n_i^{},0}]\}\hfill \end{array}$$
(18)
Separating the exponential factors in Eq.(18), we can write $`\phi _l`$ as
$$\phi _l(m,m^{};n^+,n^{})=e^{2ix(m^{}m)(p(ฯต)+i\frac{\eta }{v(ฯต)})}\stackrel{~}{\phi }_l(m,m^{};n^+,n^{})$$
(19)
with
$$\begin{array}{cc}\hfill \stackrel{~}{\phi }_l(m,m^{};n^+,n^{})=& \underset{\{m_i\}}{}\delta _{m,m_1+m_2+\mathrm{}+m_l}\underset{\{m_i^{}\}}{}\delta _{m^{},m_1^{}+m_2^{}+\mathrm{}+m_l^{}}\underset{\{n_i^+\}}{}\delta _{n^+,n_1^++n_2^++\mathrm{}+n_l^+}\underset{\{n_i^{}\}}{}\delta _{n^{},n_1^{}+n_2^{}+\mathrm{}+n_l^{}}\hfill \\ & \underset{i=1}{\overset{l}{}}\{\delta _{m_i,m_i^{}+1}\left(\genfrac{}{}{0pt}{}{m_i1+n_i^+}{m_i1}\right)\left(\genfrac{}{}{0pt}{}{m_i1+n_i^{}}{m_i1}\right)+\delta _{m_i,m_i^{}1}\left(\genfrac{}{}{0pt}{}{m_i+n_i^+}{m_i}\right)\left(\genfrac{}{}{0pt}{}{m_i+n_i^{}}{m_i}\right)\hfill \\ & +\delta _{m_i,m_i^{}}[\left(\genfrac{}{}{0pt}{}{m_i+n_i^+}{m_i}\right)\left(\genfrac{}{}{0pt}{}{m_i1+n_i^{}}{m_i1}\right)+\left(\genfrac{}{}{0pt}{}{m_i1+n_i^+}{m_i1}\right)\left(\genfrac{}{}{0pt}{}{m_i+n_i^{}}{m_i}\right)\delta _{m_i,0}\delta _{n_i^+,0}\delta _{n_i^{},0}]\}\hfill \end{array}$$
(20)
The mixing factor for the advanced GF is obtained from complex conjugation of $`\phi _{kl}`$.
Now we can express Eq.(3) in terms of $`Q`$ and the mixing factors
$$\begin{array}{cc}\hfill \rho ^k(ฯต,\varphi ;L,x)=& \left(\frac{\rho _0}{2}\right)^k\underset{l=0}{\overset{k}{}}\underset{m=0}{\overset{\mathrm{}}{}}\underset{\overline{m}=0}{\overset{\mathrm{}}{}}\underset{m^{}=0}{\overset{\mathrm{}}{}}\underset{\overline{m}^{}=0}{\overset{\mathrm{}}{}}\underset{n^+=0}{\overset{\mathrm{}}{}}\underset{n^{}=0}{\overset{\mathrm{}}{}}\underset{\overline{n^+}=0}{\overset{\mathrm{}}{}}\underset{\overline{n^{}}=0}{\overset{\mathrm{}}{}}\left(\genfrac{}{}{0pt}{}{k}{l}\right)\hfill \\ & e^{i2\pi \frac{\varphi }{\varphi _0}(n^+n^{}+\overline{n^+}\overline{n^{}})}e^{ip(ฯต)L(n^++n^{}\overline{n^+}\overline{n^{}})}e^{\frac{\eta L}{v(ฯต)}(n^++n^{}+\overline{n^+}+\overline{n^{}})}\hfill \\ & e^{2ip(ฯต)x(m^{}m\overline{m}^{}+\overline{m})}e^{\frac{2\eta x}{v(ฯต)}(m^{}m+\overline{m}^{}\overline{m})}\hfill \\ & \stackrel{~}{\phi }_l(m,m^{};n^+,n^{})\stackrel{~}{\phi }_{kl}(\overline{m},\overline{m}^{};\overline{n^+},\overline{n^{}})Q\left(\begin{array}{ccc}m,& m^{},& n^++n^{}\\ \overline{m},& \overline{m}^{},& \overline{n^+}+\overline{n^{}}\end{array}|L\right)\hfill \end{array}$$
(21)
Here, the first three exponential factors come from the external vertices and from the revolutions around the ring. The last two exponential factors were separated from the mixing coefficients \[Eq.(19)\]. Below, we shall see that the last exponential factor is canceled by a contribution from $`Q`$.
### A Equation for the central block $`Q`$
As noted above, the central block $`Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)`$ is defined as the sum of all diagrams with $`m`$ and $`\overline{m}`$ pairs of returning lines on the left side, $`m^{}`$ and $`\overline{m}^{}`$ pairs on the right side and $`n`$ and $`\overline{n}`$ through-going lines, coming from retarded and advanced GFs, respectively. An equation determining $`Q`$ can be constructed according to Berezinskiiโs idea by attaching all possible vertices, given in Fig.1, to the existing block, while avoiding the formation of unconnected electron loops. Careful analysis of all these possibilities gives the equation
$$\begin{array}{cc}\hfill \frac{\mathrm{d}}{\mathrm{d}x}& Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)=\left[\frac{1}{2l^+}(2m+n2\overline{m}\overline{n})^2+\frac{1}{l^{}}m(m+n)+\frac{1}{l^{}}\overline{m}(\overline{m}+\overline{n})+\frac{1}{2l^{}}(n+\overline{n})\right]Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)\hfill \\ & \frac{1}{l^{}}mm^{}Q\left(\begin{array}{ccc}m1,& m^{}1,& n+2\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)\frac{1}{l^{}}\overline{m}\overline{m}^{}Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m}1,& \overline{m}^{}1,& \overline{n}+2\end{array}|x\right)\hfill \\ & +\frac{1}{l^{}}m\overline{m}e^{\frac{4\eta x}{v(ฯต)}}Q\left(\begin{array}{ccc}m1,& m^{},& n\\ \overline{m}1,& \overline{m}^{},& \overline{n}\end{array}|x\right)\hfill \\ & +\frac{1}{l^{}}(m+n)(\overline{m}+\overline{n})e^{\frac{4\eta x}{v(ฯต)}}Q\left(\begin{array}{ccc}m+1,& m^{},& n\\ \overline{m}+1,& \overline{m}^{},& \overline{n}\end{array}|x\right)+\frac{1}{l^{}}m^{}(\overline{m}+\overline{n})e^{\frac{4\eta x}{v(ฯต)}}Q\left(\begin{array}{ccc}m,& m^{}1,& n+2\\ \overline{m}+1,& \overline{m}^{},& \overline{n}\end{array}|x\right)\hfill \\ & +\frac{1}{l^{}}(m+n)\overline{m}^{}e^{\frac{4\eta x}{v(ฯต)}}Q\left(\begin{array}{ccc}m+1,& m^{},& n\\ \overline{m},& \overline{m}^{}1,& \overline{n}+2\end{array}|x\right)+\frac{1}{l^{}}m^{}\overline{m}^{}e^{\frac{4\eta x}{v(ฯต)}}Q\left(\begin{array}{ccc}m,& m^{}1,& n+2\\ \overline{m},& \overline{m}^{}1,& \overline{n}+2\end{array}|x\right)\hfill \end{array}$$
(22)
The block $`Q`$ is subjected to a similar boundary condition as $`Q_0`$ in the previous section
$$Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x=0\right)=\delta _{m,0}\delta _{m^{},0}\delta _{\overline{m},0}\delta _{\overline{m}^{},0}$$
(23)
which states that for an infinitesimal ring there can be no scattering. The first coefficient on the right hand side of Eq.(22) contains contributions from the vertices a),aโ),b),bโ),c),cโ), and d) in Fig.1. Vertices a),b), and d) can be attached in $`(2m+n)`$,$`\frac{1}{2}(2m+n)(2m+n1)`$, and $`(2m+n)(2\overline{m}+\overline{n})`$ ways, respectively, the coefficients for aโ) and bโ) are the same as for a) and b), with the replacement $`\{m,n\}\{\overline{m},\overline{n}\}`$. For vertex c) we have again to distinguish two possibilities as in Sec.III. We have $`m(m+n1)`$ ways to attach it without changing $`m`$,$`m^{}`$, and $`n`$; and $`mm^{}`$ different ways with changing $`\{m,m^{}n\}\{m1,m^{}1,n+2\}`$. The latter kind of insertion of the vertex c) in Fig.1 and its counterpart for the advanced GF give the second and the third term on the right hand side of Eq.(22). The inclusion of vertex e) reduces $`m`$ and $`\overline{m}`$ by 1. The insertion of vertex f), however, can be done in four different ways which are shown schematically in the last four blocks of Fig.6.
Trying to solve Eq.(22), one may begin by substituting $`\stackrel{~}{Q}=\frac{1}{m^{}!\overline{m}^{}!}Q`$ and then introduce new variables $`M=2m+n`$, $`\overline{M}=2\overline{m}+\overline{n}`$, $`M^{}=2m^{}+n`$, $`\overline{M}^{}=2\overline{m}^{}+\overline{n}`$. As a consequence, $`M^{}`$ and $`\overline{M}^{}`$ appear as fixed parameters in the differential equation for $`\stackrel{~}{Q}`$. But still then, $`\stackrel{~}{Q}`$ depends on the five variables $`M`$,$`n`$,$`\overline{M}`$,$`\overline{n}`$, and $`x`$. Under these circumstances, looking for the general analytic solution of Eq.(22), one meets with enormous difficulties. Before studying an asymptotic approximation of the problem, we make some simplifications of Eqs.(21) and (22). The exponential factors in Eq.(22) can be removed by the following substitution:
$$Q\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)=\mathrm{exp}\left\{\frac{2\eta x}{v(ฯต)}(m+\overline{m}m^{}\overline{m}^{})\right\}\overline{Q}\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}^{},& \overline{n}\end{array}|x\right)$$
(24)
The equation for $`\overline{Q}`$ has the same structure as Eq.(22), only the exponential factors are dropped and the first term on the r.h.s. of Eq.(22) acquires another contribution $`\frac{2\eta }{v(ฯต)}(m+\overline{m}m^{}\overline{m}^{})`$ to the prefactor of $`Q`$.
From the structure of the internal vertices in Fig.1 one sees that the condition
$$(m^{}m)(\overline{m}^{}\overline{m})=0$$
(25)
is satisfied for arbitrary cross sections. (The same condition also applies to the strictly 1D problem, see \[\].) A corresponding symmetry of Eq.(22) confirms this condition.
In the regime of weak disorder, $`p_Fl1`$, Eq.(21) for the $`k`$th moment of the DOS contains terms that strongly oscillate with the particle energy ($`n^++n^{}\overline{n^+}+\overline{n^{}}`$) apart from smooth ones ($`n^++n^{}=\overline{n^+}+\overline{n^{}}`$). To neglect the strongly oscillating terms, we choose only those terms in Eq.(21) that satisfy the condition
$$n^++n^{}=\overline{n^+}+\overline{n^{}}$$
(26)
Eq.(21) now is simplified to
$$\rho ^k(ฯต,\varphi ;L)=\left(\frac{\rho _0}{2}\right)^k\underset{m=0}{\overset{\mathrm{}}{}}\underset{\overline{m}=0}{\overset{\mathrm{}}{}}\underset{m^{}=0}{\overset{\mathrm{}}{}}\underset{n=0}{\overset{\mathrm{}}{}}e^{i4\pi \frac{\varphi }{\varphi _0}n\frac{2\eta L}{v(ฯต)}n}\overline{Q}\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}m+m^{},& n\end{array}|x\right)\mathrm{\Phi }_k(m,\overline{m},m^{},\overline{m}m+m^{},n,n)$$
(27)
where
$$\mathrm{\Phi }_k(m,\overline{m},m^{},\overline{m}^{},n,\overline{n})=\underset{l=0}{\overset{k}{}}\underset{n^+=0}{\overset{n}{}}\underset{\overline{n^+}=0}{\overset{\overline{n}}{}}\left(\genfrac{}{}{0pt}{}{k}{l}\right)e^{i4\pi \frac{\varphi }{\varphi _0}(n^++\overline{n^+})}\stackrel{~}{\phi }_l(m,m^{};n^+,nn^+)\stackrel{~}{\phi }_{kl}(\overline{m},\overline{m}^{};\overline{n^+},\overline{n}\overline{n^+})$$
(28)
Expressions (27) and (28) show that the dominating contribution to $`\rho ^k`$ does not strongly oscillate with the energy. Unlike the averaged DOS, the first harmonic of all moments oscillates with the halved magnetic flux $`\frac{\varphi _0}{2}`$.
In the following section, we solve Eqs.(20)-(28) for large rings, with $`L\mathrm{max}\{l^+,l^{}\}`$.
## V Distribution functions for the weak localization limit
The equation for $`\overline{Q}`$ is simplified considerably in the limit of large rings, $`L\mathrm{max}\{l^+,l^{}\}`$. For this case, we can assume that the electrons are quasiโlocalized and that the wave function overlaps around the ring are small, similar to a tightโbinding model. Diagrammatically, this means that the electron loops emerging from the l.h.s. and the r.h.s. of the diagram almost never reach each other, since they have a characteristic size of $`\xi L`$. (The localization length for an infinite 1D system is $`\xi _{1D}4l^{}`$.)
As a consequence, we can for large rings neglect those inclusions of the vertices c),cโ),e), and f) that directly connect the loops on the r.h.s. with those on the l.h.s.. Corresponding to this is the neglect of the terms 2,3, and 6-10 on the r.h.s of Eq.(22). Now, $`\overline{Q}`$ can be factored as
$$\overline{Q}\left(\begin{array}{ccc}m,& m^{},& n\\ \overline{m},& \overline{m}m+m^{},& n\end{array}|x\right)=Q^{}(m,\overline{m},n;x)Q^{}(m^{},\overline{m}m+m^{},n;x)$$
(29)
where the factors are defined through
$$\begin{array}{cc}\hfill \frac{\mathrm{d}Q^{}(m,\overline{m},n;x)}{\mathrm{d}x}=& \left[\frac{2\eta }{v(ฯต)}(m+\overline{m})+\frac{2}{l^+}(m\overline{m})^2+\frac{1}{l^{}}m(m+n)+\frac{1}{l^{}}\overline{m}(\overline{m}+\overline{n})+\frac{1}{l^{}}n\right]Q^{}(m,\overline{m},n;x)\hfill \\ & +\frac{1}{l^{}}m\overline{m}Q^{}(m1,\overline{m}1,n;x)+\frac{1}{l^{}}(m+n)(\overline{m}+\overline{n})Q^{}(m+1,\overline{m}+1,n;x)\hfill \end{array}$$
(30)
Apart from this simplification, the limit $`L\mathrm{max}(l^+,l^{})`$ implies $`m,\overline{m},m^{}n`$.
Note that Eq.(30) can also be obtained from Eqs.(22) and (24) by neglecting $`m^{}`$ and $`\overline{m}^{}`$. The non-entanglement mentioned above has a second consequence: The remaining contributions change $`m`$ and $`\overline{m}`$ simultaneously by $`\pm 1`$ (due to vertices e) and f) of Fig.1), or conserve both $`m`$ and $`\overline{m}`$, as in Berezinskiiโs approach to strictly 1D systems. Therefore we can adopt $`m=\overline{m}`$, which further simplifies Eq.(30):
$$\begin{array}{cc}\hfill l^{}\frac{\mathrm{d}Q^{}(m,n;x)}{\mathrm{d}x}=& \left[4\eta \tau ^{}m+n+2m(m+n)\right]Q^{}(m,n;x)+m^2Q^{}(m1,n;x)\hfill \\ & +(m+n)^2Q^{}(m+1,n;x)\hfill \end{array}$$
(31)
Here, $`\tau ^{}`$ is the inelastic scattering time with respect to backward scattering. The boundary condition for Eq.(31) is
$$Q^{}(m,n;x=0)=\delta _{m,0}$$
(32)
The combined mixing function $`\mathrm{\Phi }_k`$, given in Eq.(28) is simplified for large $`m`$, $`\overline{m}`$, and $`m^{}`$ in Appendix B. Substituting Eqs.(B6) and (29) into (27), we get a comparatively simple expression for $`\rho ^k`$:
$$\begin{array}{cc}\hfill \rho ^k(ฯต,\varphi ;L)=& \left(\frac{\rho _0}{2}\right)^k\underset{m=0}{\overset{\mathrm{}}{}}\underset{m^{}=0}{\overset{\mathrm{}}{}}\underset{n=0}{\overset{\mathrm{}}{}}\underset{l=0}{\overset{k}{}}\left[\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0})\right]^{2n}e^{\frac{2\eta L}{v(ฯต)}n}\left(\genfrac{}{}{0pt}{}{k}{l}\right)\left(\genfrac{}{}{0pt}{}{2l}{mm^{}}\right)\left(\genfrac{}{}{0pt}{}{2k2l}{mm^{}}\right)\hfill \\ & \frac{2^{2n}m^{k+2n2}}{\mathrm{\Gamma }(l)\mathrm{\Gamma }(kl)}Q^{}(m,n;L)Q^{}(m^{},n;L)\hfill \end{array}$$
(33)
As we emphasized, the diagrammatical structure of the block $`Q^{}`$ demands its dependence on one parameter $`m`$ instead of two ($`m`$ and $`\overline{m}`$). Hereby the sum over $`\overline{m}`$ is removed. The summations over $`l`$ and $`m^{}`$ can be done as described in Appendix B. Using Eqs. (B7) and (B8), we get
$$\begin{array}{cc}\hfill \rho ^k(ฯต,\varphi ;L)& =\rho _0^k\frac{2^{1k}(k1)}{k(2k1)}\frac{\mathrm{\Gamma }^2(2k)}{\mathrm{\Gamma }^5(k)}\underset{n=0}{\overset{\mathrm{}}{}}\underset{m=0}{\overset{\mathrm{}}{}}\frac{2^{2n}e^{\frac{2\eta L}{v(ฯต)}n}m^{k+2n2}}{\mathrm{\Gamma }^2(n+1)}\left[\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0})\right]^{2n}Q_{}^{}{}_{}{}^{2}(m,n;L)\hfill \\ & =\rho ^k(ฯต,\varphi ;L)_{n=0}+\rho ^k(ฯต,\varphi ;L)_{n=1}\mathrm{cos}^2(\frac{2\pi \varphi }{\varphi _0})+\mathrm{}\hfill \end{array}$$
(34)
Eq.(31) for $`Q^{}(m,n;x)`$ was solved approximately in \[\] for arbitrary $`n`$. Here we shall study this equation for the zeroth and first harmonics ($`n=0`$ and $`n=1`$) in detail.
### A Zeroth harmonic contribution to the DOS moments
The zeroth harmonic of $`\rho ^k(ฯต,\varphi ;L)`$ in Eq.(34) contains $`Q^{}(m,n=0;L)`$. Laplace transforming Eq.(31), written for $`n=0`$, with respect to $`x`$ and using the boundary condition (32), we get
$$(\lambda +s_1m)Q_0^{}(m;\lambda )\delta _{m,0}=m^2\left[Q_0^{}(m+1;\lambda )+Q_0^{}(m1;\lambda )2Q_0^{}(m;\lambda )\right]$$
(35)
Here, $`s_1=4\eta \tau ^{}`$ and $`\lambda `$ is the parameter of the Laplace transform. This is an equation for the right hand side in the Berezinskii technique with an open boundary condition and it was solved in \[\]. Here, we give only the result for $`Q_0^{}(s_1,m;x)`$:
$$Q_0^{}(s_1,m;x)=2(ms_1)^{\frac{1}{2}}\mathrm{K}_1(2(ms_1)^{\frac{1}{2}})+\frac{2(ms_1)^{\frac{1}{2}}}{2\pi i}_{\mathrm{}}^{\mathrm{}}d\lambda \frac{s_1^{\frac{1+i\lambda }{2}}}{i\lambda }\frac{\mathrm{\Gamma }^3(\frac{1i\lambda }{2})}{\mathrm{\Gamma }^2(i\lambda )}e^{(\lambda ^2+1)\frac{x}{4l^{}}}\mathrm{K}_{i\lambda }(2(ms_1)^{\frac{1}{2}})$$
(36)
After substitution of this solution into $`\rho ^k(ฯต,\varphi ;L)_{n=0}`$ in Eq.(34), the summation over $`m`$ can be transformed into an integration, which is done easier. Some mathematics results in the following form of $`\rho ^k(ฯต,\varphi ;L)_{n=0}`$:
$$\begin{array}{cc}\hfill \rho ^k(ฯต,\varphi ;L)_{n=0}=& \left(\frac{\rho _0}{2}\right)^k\frac{2(k1)\mathrm{\Gamma }(2k)}{k(2k1)\mathrm{\Gamma }^5(k)}s_1^{1k}\{\frac{k}{k1}\mathrm{\Gamma }^4(k)\hfill \\ & +\frac{2}{\sqrt{s_1}}e^{\frac{L}{4}}_{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}\lambda }{2\pi i}\frac{e^{\frac{L}{4l^{}}(\lambda i\gamma )^2\frac{L}{4l^{}}\gamma ^2}}{i\lambda }\frac{\mathrm{\Gamma }^3\left(\frac{1i\lambda }{2}\right)}{\mathrm{\Gamma }^2(i\lambda )}\left|\mathrm{\Gamma }\left(\frac{2k+1+i\lambda }{2}\right)\mathrm{\Gamma }\left(\frac{2k1+i\lambda }{2}\right)\right|^2\hfill \\ & +\frac{2}{s_1}_{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}z^{}}{2\pi i}e^{\frac{L}{2l^{}}z_{}^{}{}_{}{}^{2}}|\mathrm{\Gamma }(kiz^{})|^2\hfill \\ & _{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}z}{2\pi i}\frac{e^{\frac{L}{2l^{}}(zi\gamma )^2\frac{L}{2l^{}}\gamma ^2\frac{L}{2l^{}}}}{(z+z^{}i)(zz^{}i)}\frac{\mathrm{\Gamma }^3\left(\frac{1iziz^{}}{2}\right)\mathrm{\Gamma }^3\left(\frac{1iz+iz^{}}{2}\right)}{\mathrm{\Gamma }^2(iziz^{})\mathrm{\Gamma }^2(iz+iz^{})}|\mathrm{\Gamma }(kiz)|^2\}\hfill \end{array}$$
(37)
where $`\gamma =\frac{l^{}}{L}\mathrm{ln}\frac{1}{s_1}>0`$. The second term in the bracket of Eq.(37) has a saddle point at $`\lambda _0=i\gamma `$ and simple poles at the upper halfโplane: $`\lambda _1=i,\lambda _2=i(2k1),i(2k+1),\mathrm{}`$. The integral over $`z`$ in the third term in the bracket contains again the saddle point at $`z_0=i\gamma `$ and poles at $`z_1=\pm z^{}+i,z_2=ik,i(k+1),\mathrm{}`$. For $`\gamma <1`$, the main contribution to both integrals is given by the saddle points. As a result we get
$$\rho ^k(ฯต,\varphi ;L)_{n=0}=\rho _0^k(2s_1)^{1k}\frac{\mathrm{\Gamma }(2k1)}{\mathrm{\Gamma }(k)}$$
(38)
Such a result has been obtained for the infinite 1D disordered system. Transforming the semiโinvariants in Eq.(38) to moments and using the inverse Mellin transformation
$$W(\rho )=\frac{1}{2\pi i}_{ai\mathrm{}}^{a+i\mathrm{}}\frac{\mathrm{d}k}{\rho ^{k+1}}\rho ^k$$
(39)
the following inverseโGaussian distribution function is obtained:
$$W_{n=0}(\rho )=\left(\frac{2\eta \tau ^{}\rho _0}{\pi \rho ^3}\right)^{\frac{1}{2}}\mathrm{exp}\left(2\frac{(\rho \rho _0)^2}{\rho \rho _0}\eta \tau ^{}\right)$$
(40)
For $`2\eta \tau ^{}1`$, the most probable or typical value of $`\rho `$ is equal to $`\rho _0`$, whereas for $`2\eta \tau ^{}1`$ it shifts to lower values and becomes equal to $`\rho _{\mathrm{typ}}=\frac{4\eta \tau ^{}}{3}\rho _0`$.
When $`\gamma `$ assumes intermediate values, i.e. $`1<\gamma <k`$, the essential contribution to $`\rho ^k_{n=0}`$ comes from the saddle points of the third term in the bracket of Eq.(37) and the contributions from the poles of this term cancel the other term, resulting in
$$\rho ^k(ฯต,\varphi ;L)_{n=0}=\frac{\rho _0^k\mathrm{\Gamma }(2k1)}{2^{k1}\mathrm{\Gamma }(k1)\mathrm{\Gamma }(k+1)}\frac{s_1^kl^{}e^{\frac{L}{2l^{}}(1+\gamma ^2)}}{\pi L(1\gamma )^2}\frac{\mathrm{\Gamma }^6\left(\frac{1+\gamma }{2}\right)}{\mathrm{\Gamma }^4(\gamma )}\mathrm{\Gamma }(k+\gamma )\mathrm{\Gamma }(k\gamma )$$
(41)
This expression shows that high moments of the DOS for intermediate values of $`\gamma `$ increase with $`k`$; however the increase is not so rapid, Eq.(41) has an additional factor $`\frac{1}{k!}`$ compared to Eq.(38).
For $`\gamma `$ satisfying the condition $`\gamma >k`$, the leading contribution is given by the pole $`z_2=ik`$ and
$$\rho ^k(ฯต,\varphi ;L)_{n=0}=\left(\frac{\rho _0}{2}\right)^k\sqrt{\frac{l^{}}{2\pi L}}\frac{4}{k(k1)(2k1)}\frac{\mathrm{\Gamma }^2(2k)}{\mathrm{\Gamma }^7(k)}\mathrm{\Gamma }^6\left(\frac{k+1}{2}\right)e^{\frac{L}{2l^{}}(k^21)}$$
(42)
The last expression is valid for arbitrary small values of the dissipation parameter ($`\eta 0`$ or $`\gamma \mathrm{}`$) with $`\eta \frac{1}{4\tau ^{}}\mathrm{exp}(\frac{kL}{l^{}})`$. Eq.(42) shows that the zeroth harmonic of the $`k`$th moment of the DOS grows with $`k`$ as $`\mathrm{exp}(k^2)`$. Such rapid increasing of high moments of $`\rho ^k_{n=0}`$ has been firstly obtained by Wegner and it is a characteristic feature of the logarithmic normal distribution of $`\rho ^k_{n=0}`$. The distribution function for the zeroth harmonic term can be obtained using Eq.(39). For large values of the DOS, satisfying the condition $`\rho >\frac{\rho _0}{2}\mathrm{exp}(\frac{L}{l^{}})`$, the dominating saddle point yields again a logarithmic normal distribution:
$$W_{n=0}(\rho )=\frac{8l^{}}{\pi \rho _0L}\frac{\mathrm{\Gamma }\left(\frac{2l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}2\right)\mathrm{\Gamma }\left(\frac{2l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)\mathrm{\Gamma }^6\left(\frac{1}{2}+\frac{l^{}}{2L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)}{\mathrm{\Gamma }\left(\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}+1\right)\mathrm{\Gamma }^6\left(\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)}\mathrm{exp}\left[\frac{l^{}}{2L}\left(\mathrm{ln}\frac{2\rho }{\rho _0}+\frac{L}{l^{}}\right)^2\right]$$
(43)
For small values of $`\rho `$, when $`\rho <\frac{\rho _0}{2}\mathrm{exp}(\frac{L}{l^{}})`$, the main contribution comes from the pole at the origin and the distribution function decreases powerโlike:
$$W_{n=0}(\rho )=\frac{2\rho _0}{\rho ^2}\sqrt{\frac{l^{}}{2\pi L}}$$
(44)
Thus the distribution function for the zeroth harmonic or $`\varphi `$-independent component has asymmetric form.
### B Amplitude of the first harmonic contribution to the DOS moments
By a Laplace transform with respect to $`x`$, Eq.(31) with $`n=1`$ is converted to
$$(\lambda +s_1m)Q_1^{}(m;\lambda )\delta _{m,0}=(m+1)^2[Q_1^{}(m+1;\lambda )Q_1^{}(m;\lambda )]+m^2[Q_1^{}(m1;\lambda )Q_1^{}(m;\lambda )]$$
(45)
The $`\delta `$ symbol on the leftโhand side of this equation comes from the boundary condition (32). Eq.(45) corresponds to the equation for the central part in the Berezinskii technique for strictly 1D systems with open boundary. For $`m1`$, this equation is transformed into a differential equation,
$$m^2\frac{\mathrm{d}^2Q_1^{}(m;\lambda )}{\mathrm{d}m^2}+2m\frac{\mathrm{d}Q_1^{}(m;\lambda )}{\mathrm{d}m}(\lambda +s_1m)Q_1^{}(m;\lambda )=0$$
(46)
A change of the function to $`\frac{1}{z}\mathrm{\Phi }(z,\lambda )=Q_1^{}(m;\lambda )`$, where $`z^2=4ms_1`$, reduces Eq.(46) to the Bessel equation
$$\frac{\mathrm{d}^2\mathrm{\Phi }}{\mathrm{d}z^2}+\frac{1}{z}\frac{\mathrm{d}\mathrm{\Phi }}{\mathrm{d}z}\left(1+\frac{1+4\lambda }{z^2}\right)\mathrm{\Phi }=0$$
(47)
Therefore $`Q_1^{}(m;\lambda )`$ can be expressed as
$$Q_1^{}(m;\lambda )=C\frac{1}{2(ms_1)^{\frac{1}{2}}}\mathrm{K}_{1+2q}(2(ms_1)^{\frac{1}{2}})$$
(48)
where $`q=\frac{1}{2}+\sqrt{\lambda +\frac{1}{4}}`$.
Eq.(48) contains an unknown parameter $`C`$ due to the neglect of the Kronecker symbol in Eq.(45).
On the other hand Eq.(45) has been solved by Melโnikov, who obtained the asymptotic solution of $`Q_1^{}(m;\lambda )`$ for $`1ms_1^1`$ as
$$Q_1^{}(m;\lambda )=\frac{\mathrm{\Gamma }^3(q+1)}{\mathrm{\Gamma }(2q+2)}m^{q1}$$
(49)
The comparison or Eq.(48) with the asymptotic form (49) allows to determine $`C`$:
$$C=4s_1^{q+1}\frac{\mathrm{\Gamma }^3(q+1)}{\mathrm{\Gamma }(2q+2)\mathrm{\Gamma }(2q+1)}$$
(50)
Taking the inverse Laplace transform, one obtains for $`Q_1^{}(m;x)`$
$$Q_1^{}(m;x)=_{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}\lambda }{2\pi }e^{\frac{1}{4}(\lambda ^2+1)\frac{x}{l^{}}}s_1^{\frac{1i\lambda }{2}}(ms_1)^{\frac{1}{2}}\mathrm{K}_{i\lambda }(2(ms_1)^{\frac{1}{2}})\frac{\mathrm{\Gamma }^3\left(\frac{1i\lambda }{2}\right)}{\mathrm{\Gamma }^2(i\lambda )}$$
(51)
To get an expression for the first harmonic, $`\rho ^k(ฯต,L)_{n=1}`$, we substitute the solution (51) into Eq.(34), and sum over $`m`$, which can be done after the transformation of the sum into an integral over $`\kappa =ms_1`$:
$$\begin{array}{cc}\hfill \rho ^k(ฯต,L)_{n=1}=& \left(\frac{\rho _0}{2}\right)^k\frac{(k1)\mathrm{\Gamma }(2k)}{k\mathrm{\Gamma }^5(k)}s_1^ke^{\frac{L}{2l^{}}}_{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}\lambda }{2\pi }s_1^{\frac{i\lambda }{2}}e^{\frac{L}{4l^{}}\lambda ^2}\frac{\mathrm{\Gamma }^3\left(\frac{1i\lambda }{2}\right)}{\mathrm{\Gamma }^2(i\lambda )}\hfill \\ & _{\mathrm{}}^{\mathrm{}}\frac{\mathrm{d}\lambda ^{}}{2\pi }s_1^{\frac{i\lambda ^{}}{2}}e^{\frac{L}{4l^{}}\lambda _{}^{}{}_{}{}^{2}}\frac{\mathrm{\Gamma }^3\left(\frac{1i\lambda ^{}}{2}\right)}{\mathrm{\Gamma }^2(i\lambda ^{})}\left|\mathrm{\Gamma }\left(k+\frac{i\lambda +i\lambda ^{}1}{2}\right)\mathrm{\Gamma }\left(k+\frac{i\lambda i\lambda ^{}1}{2}\right)\right|^2\hfill \end{array}$$
(52)
For convenience, we substitute below the variables $`\lambda `$ and $`\lambda ^{}`$ by $`z`$ and $`z^{}`$ according to $`\lambda =z+z^{}`$ and $`\lambda ^{}=zz^{}`$.
The values of the integrals in Eq.(52) are determined by saddle points and poles. For $`\gamma =\frac{l^{}}{L}\mathrm{ln}\frac{1}{s_1}<k\frac{1}{2}`$, the contribution from the saddle point dominates:
$$\rho ^k(ฯต,L)_{n=1}=\left(\frac{\rho _0}{2}\right)^k\frac{(k1)l^{}\mathrm{\Gamma }^2(2k)s_1^k}{k\pi L\mathrm{\Gamma }^5(k)}e^{\frac{L}{2l^{}}(1+\gamma ^2)}\frac{\mathrm{\Gamma }^2\left(\frac{2k1}{2}\right)\mathrm{\Gamma }^6\left(\frac{1+\gamma }{2}\right)\mathrm{\Gamma }\left(\frac{2k1+2\gamma }{2}\right)\mathrm{\Gamma }\left(\frac{2k12\gamma }{2}\right)}{\mathrm{\Gamma }^4(\gamma )}$$
(53)
For $`\gamma >k\frac{1}{2}`$ the main contribution is given by the pole at $`z=i(k\frac{1}{2})`$ and one gets
$$\rho ^k(ฯต,L)_{n=1}=\left(\frac{\rho _0}{2}\right)^k\frac{2(k1)\mathrm{\Gamma }(2k)\mathrm{\Gamma }(2k1)\mathrm{\Gamma }^6\left(\frac{2k+1}{4}\right)}{k\mathrm{\Gamma }^5(k)\mathrm{\Gamma }^2\left(\frac{2k1}{2}\right)}\sqrt{\frac{l^{}}{2\pi Ls_1}}e^{\frac{L}{8l^{}}(2k1)^2\frac{L}{2l^{}}}$$
(54)
In contrast to the expression of $`\rho ^k(ฯต,L)_{n=0}`$ for small dissipation, $`\eta 0`$ \[see Eq.(42)\], the expression for $`\rho ^k(ฯต,L)_{n=1}`$ increases strongly with $`s_1=4\tau ^{}\eta 0`$. It is illustrative to rewrite the prefactor of Eq.(54) as $`(2\pi \frac{L}{l^{}}s_1)^{\frac{1}{2}}=(8\pi ^2\frac{\eta }{\mathrm{\Delta }})^{\frac{1}{2}}`$ in terms of the level distance $`\mathrm{\Delta }=\frac{1}{\rho _0L}`$ and the dissipation energy $`\eta `$, the latter blurring the quantized energy levels. By decreasing $`\eta `$, the energy levels are sharpened and the distribution function becomes a $`\delta `$โfunction.
Substituting Eq.(54) into (39), one receives a normal logarithmic distribution for $`\rho >\frac{\rho _0}{2}\mathrm{exp}(\frac{L}{l^{}})`$:
$$W_{n=1}(\rho )=\frac{l^{}}{\pi L}\sqrt{\frac{2}{s_1\rho \rho _0}}\frac{\mathrm{\Gamma }\left(1+\frac{2l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)\mathrm{\Gamma }\left(\frac{2l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)\mathrm{\Gamma }^6\left(\frac{1}{2}+\frac{l^{}}{2L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)}{\mathrm{\Gamma }\left(\frac{3}{2}+\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)\mathrm{\Gamma }\left(\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\frac{1}{2}\right)\mathrm{\Gamma }^3\left(\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}+\frac{1}{2}\right)\mathrm{\Gamma }^2\left(\frac{l^{}}{L}\mathrm{ln}\frac{2\rho }{\rho _0}\right)}\mathrm{exp}\left[\frac{l^{}}{2L}\left(\mathrm{ln}\frac{2\rho }{\rho _0}+\frac{L}{l^{}}\right)^2\right]$$
(55)
The logarithmic normal distribution function for the first harmonic is valid for a large range of $`\rho `$. Comparing Eq.(55) with Eq.(43) for winding number zero, it can be seen that Eq.(55) contains in addition a prefactor $`\sqrt{\frac{\rho _0}{\eta \tau ^{}\rho }}=(\pi \eta l^{}\rho )^{1/2}`$ which increases with decreasing temperature. Thus the first harmonic increases with decreasing temperature faster in amplitude than the zeroth harmonic.
## VI Conclusion
The distribution function for the local DOS in a oneโchannel ring threaded by a magnetic flux through the opening was studied in this paper. For this purpose, we constructed a new diagrammatic method as an extension of the Berezinskii technique to the problem with periodic boundary conditions and in the presence of an external magnetic field. The equations obtained ((10) to (12) and (21) to (23) for the DOS and its $`k`$th moments, respectively) are exact in the framework of the weak disorder limit $`k_Fl1`$. Eqs.(11) and (12) are solved exactly, which gives the oscillation of $`\rho `$ with the full flux for both weak localization and ballistic regimes.
In contrast to the DOS problem, the equation for $`\rho ^k(ฯต,\varphi ;L)`$ is rather complicated and we succeeded to solve it for the weak localization limit when $`Ll^\pm `$. In this limit, the leading contributions to arbitrary moments of the DOS oscillate with the halved period $`\frac{\varphi _0}{2}`$. The distribution functions for zeroth (insensitive to the magnetic field) and first (with a period of $`\frac{\varphi _0}{2}`$) harmonics are calculated and logarithmic normal distributions (Eqs.(43) and (55)) are obtained for them, indicating large contributions from high moments of the DOS. For the zeroth harmonic, this normal logarithmic shape appears for the tail of the distribution, but for the first harmonic it covers the large range of $`\rho >\frac{\rho _0}{2}\mathrm{exp}(\frac{L}{l^{}})`$, i.e. the high moments give essential contributions not only on the tail but also in the vicinity of the average value of the DOS. The distribution function for the first harmonic increases with decreasing the width of the energy levels or the dissipation parameter $`\eta `$ (see Eq.(55)), which was introduced phenomenologically in the theory (Eq.(4)). For $`\eta 0`$, the distribution function $`W_{n=1}(\rho )`$ becomes a $`\delta `$-function due to the quantization of the energy levels in the rings. The results for the DOS show that the amplitudes of all harmonics of $`\rho (ฯต,\varphi )`$ are exponentially small in the weak localization regime \[Eq.(14)\], while the amplitudes of the higher moments in this regime \[Eqs.(38), (41), (42), and (54)\] are relatively large. Although we could not calculate higher moments of the DOS in the ballistic regime, the amplitude of the average value of the DOS is large and seems to be consistent with experimental data.
It is also well known that the DOS of 1D disordered crystalline systems is very sensitive to the filling factor. There exists disorder induced enhancement of the DOS for commensurable values of the electron wavelength $`\lambda `$ and the lattice constant $`a`$, when the electronic energy $`ฯต`$ satisfies the condition $`p(ฯต)=\frac{k\pi }{na}`$, $`k=\pm 1,\pm 2,\mathrm{}\pm n`$ and $`n=2,3,\mathrm{}`$, and the effect is pronounced for halfโfilling which corresponds to $`n=2`$. The singularity in the DOS of 1D disordered crystalline systems near the middle of the band is known as a Dyson singularity which was studied for many 1D electronic models. Notice that the Berezinskii method has also been applied to study the conductivity and the localization length apart from the Dyson singularity in the middle of the band of a 1D infinite lattice with both weak and strong disorder. Our preliminary study shows that the real space diagrammatic method presented in this paper is applicable to study the Dyson singularity in the DOS of a ring for a halfโfilled energy band. This leads to a remarkable high amplitude of the persistent current as it is observed in the experiments, provided that the Peierls transition is suppressed by impurities and by weak transvers tunneling between the channels.
The authors thank M. Kiselev for discussion. This work was supported by the SFB410.
## A Solution for $`Q_0(m,n;x)`$
The first term on the right hand side of Eq.(11) can be removed through the transformation
$$Q_0(m,n;x)=\mathrm{exp}\left\{\frac{x}{2l^+}(2m+n)^2\frac{x}{2l^{}}n\frac{x}{l^{}}m(m+n)\right\}Q_0^{}(m,n;x)$$
(A1)
which gives for Eqs.(11) and (12) the simpler form
$$l^{}\frac{\mathrm{d}Q_0^{}(m,n;x)}{\mathrm{d}x}=m^2\mathrm{exp}\left\{\frac{xn}{l^{}}\right\}Q_0^{}(m1,n+2;x)\text{with}Q_0^{}(m,n;x=0)=\delta _{m,0}$$
(A2)
Laplace transformation from $`x`$ to $`\lambda `$ yields
$$\lambda \overline{Q_0}(m,n;\lambda )\delta _{m,0}=\frac{1}{l^{}}m^2\overline{Q_0}(m1,n+2;x\frac{n}{l^{}})$$
(A3)
Eq.(A3) can be solved by iteration:
$$\begin{array}{c}\hfill \overline{Q_0}(m,n;\lambda )=\left(\frac{1}{l^{}}\right)^m(m!)^2\underset{j=0}{\overset{m}{}}\frac{1}{\lambda \frac{1}{l^{}}j(j+n1)}\\ \hfill =\frac{(m!)^2}{\lambda }\frac{\mathrm{\Gamma }(z+1+\frac{n1}{2})\mathrm{\Gamma }(\frac{n1}{2}+1z)}{\mathrm{\Gamma }(\frac{n1}{2}+m+1+z)\mathrm{\Gamma }(\frac{n1}{2}+m+1z)}\end{array}$$
(A4)
where $`z^2=\lambda l^{}+\frac{(n1)^2}{4}`$. The inverse Laplace transform gives for $`Q_0^{}`$
$$Q_0^{}(m,n;x)=(m!)^2\underset{j=0}{\overset{m}{}}\mathrm{exp}\left\{\frac{x}{l^{}}j(j+n1)\right\}\frac{(1)^j}{j!(mj)!}\frac{(j+n2)!}{(m+j+n1)!}(2j+n1)$$
(A5)
which, in connection with (A1), gives the final result Eq.(13), where $`x`$ is replaced by the full circumference $`L`$. The compliance of Eq.(A5) with the boundary condition is easily checked. Also, for $`m=0`$, $`Q_0^{}(0,n;x)=1`$, and for $`n=0`$ we get from the inverse Laplace transform of Eq.(A4) or from taking the limit of Eq.(A5)
$$Q_0^{}(m,0;x)=(1mm\frac{x}{l^{}})+\underset{j=2}{\overset{m}{}}\mathrm{exp}\left\{\frac{x}{l^{}}j(j+n1)\right\}\left(\genfrac{}{}{0pt}{}{m}{j}\right)(1)^j\frac{m!(j+n2)!}{(m+j+n1)!}(2j+n1)$$
(A6)
## B Calculation of the mixing coefficient
Using the relations
$$\delta _{m,k}=_{|z|<1}\frac{\mathrm{d}z}{2\pi i}z^{km1}\text{and}\left(\genfrac{}{}{0pt}{}{m}{k}\right)=_{|z|<1}\frac{\mathrm{d}z}{2\pi i}\frac{1}{z^{k+1}(1z)^{mk+1}}$$
(B1)
we can transform Eq.(20) for $`\stackrel{~}{\phi }_l(m,m^{};n^+,nn^+)`$ to
$$\stackrel{~}{\phi }_l(m,m;n^+,nn^+)=\frac{\mathrm{d}z_1}{2\pi i}\frac{1}{z_1^{m+1}}\frac{\mathrm{d}z_2}{2\pi i}\frac{1}{z_2^{m^{}+1}}\frac{\mathrm{d}z_3}{2\pi i}\frac{1}{z_3^{n^++1}}\frac{\mathrm{d}z_4}{2\pi i}\frac{1}{z_4^{nn^++1}}\left\{\frac{(1+z_1)(1+z_2)z_3z_4}{(1z_3)(1z_4)z_1z_2}\right\}^l$$
(B2)
Substituting $`z=z_1z_2`$, the dominant contribution for $`m1`$ comes from the pole at $`z=(1z_3)(1z_4)`$. Integrating over this new variable gives
$$\begin{array}{cc}\hfill \stackrel{~}{\phi }_l(m,m^{};n^+,nn^+)=& \frac{(m+l1)!}{(l1)!m!}\frac{\mathrm{d}z_2}{2\pi i}\frac{1}{z_2^{m^{}m+l+1}}\frac{\mathrm{d}z_3}{2\pi i}\frac{1}{z_3^{n^++1}}\frac{\mathrm{d}z_4}{2\pi i}\frac{1}{z_4^{nn^++1}}\hfill \\ & \frac{\left[(z_2+(1z_3)(1z_4))(1+z_2)z_2z_3z_4\right]^l}{(1z_3)^{m+l}(1z_4)^{m+l}}\hfill \end{array}$$
(B3)
The remaining integrals are done in a similar way, resulting in
$$\stackrel{~}{\phi }_l(m,m^{};n^+,nn^+)=\frac{(2l)!(m+l+n^+1)!(m+l+nn^+1)!}{(m^{}m+l)!(mm^{}+l)!(m+l1)!n^+!(nn^+)!(l1)!m!}$$
(B4)
Now we can collect all $`n^+`$ and $`\overline{n^+}`$ dependent terms in Eq.(21) and sum over $`n^+`$ and $`\overline{n^+}`$, introducing the mixing function $`\mathrm{\Phi }_k`$ from Eq.(28). For large $`m`$, we can use Stirlings formula
$$\underset{m\mathrm{}}{lim}\frac{(m+a)!}{(m+b)!}m^{ba}=1$$
(B5)
to obtain
$$\mathrm{\Phi }_k(m,\overline{m},m^{},\overline{m}^{},n,\overline{n})=\underset{l=0}{\overset{k}{}}\frac{\left(1+e^{4i\pi \frac{\varphi }{\varphi _0}}\right)^{n+\overline{n}}}{(l1)!(kl1)!n!\overline{n}!}\left(\genfrac{}{}{0pt}{}{k}{l}\right)\left(\genfrac{}{}{0pt}{}{2l}{mm^{}+l}\right)\left(\genfrac{}{}{0pt}{}{2k2l}{\overline{m}\overline{m}^{}+l}\right)$$
(B6)
Taking into account $`Q^{}(m,\overline{m},n;x)=Q^{}(m,n;x)\delta _{m,\overline{m}}`$, this gives Eq.(33). For $`m,m^{}1`$, the summations over $`l`$ and $`m^{}`$ can be done. Following the paper \[\], we denote $`\mathrm{\Delta }m=mm^{}`$, with $`\mathrm{\Delta }mm`$ for large $`m`$. The significant contributions to Eq.(33) come from $`Q^{}(m\mathrm{\Delta }m,n;x)Q^{}(m,n;x)`$. Hence we can rewrite Eq.(33) as
$$\begin{array}{cc}\hfill \rho ^k(ฯต,\varphi ;L)=& \left(\frac{\rho _0}{2}\right)^k\underset{m=0}{\overset{\mathrm{}}{}}\underset{n=0}{\overset{\mathrm{}}{}}\left[\mathrm{cos}(2\pi \frac{\varphi }{\varphi _0})\right]^{2n}e^{\frac{2\eta L}{v(ฯต)}n}\frac{2^{2n}m^{k+2n2}k!}{(n!)^2}Q_{}^{}{}_{}{}^{2}(m,n;x)\hfill \\ & \underset{l=0}{\overset{k}{}}\left(\genfrac{}{}{0pt}{}{k}{l}\right)\underset{\mathrm{\Delta }m=k}{\overset{k}{}}\left(\genfrac{}{}{0pt}{}{2l}{l+\mathrm{\Delta }m}\right)\left(\genfrac{}{}{0pt}{}{2k2l}{kl+\mathrm{\Delta }m}\right)\frac{1}{(l1)!(kl1)!}\hfill \end{array}$$
(B7)
The last two sums result in
$$\frac{2(k1)}{k(2k1)}\frac{\mathrm{\Gamma }^2(2k)}{\mathrm{\Gamma }^5(k)}$$
(B8)
which is used in the final result for the $`k`$th moment of the DOS, Eq.(34). |
warning/0003/math0003128.html | ar5iv | text | # Quantum Lefschetz Hyperplane Theorem
## 1. Introduction
### 1.1. Statement of the main results
Let $`X`$ be a smooth projective variety. Let $`T_k`$ be a basis of $`H^{}(X)`$, $`T_0=1,T_i=p_i,i=1,\mathrm{}N`$ and let $`t_k`$ be the dual coordinates of $`T_k`$. Let $`g_{mk}:=_XT_mT_k`$ the intersection pairing and $`(g^{mk})`$ is the inverse matrix of $`(g_{mk})`$. Let $`E:=(_jL_j)`$ be a vector bundle on $`X`$, where $`L_j`$ are pull-backs of convex/concave (see below) line bundles on $`P`$. Define the generating function of genus zero one-point gravitational GromovโWitten invariants on $`X`$ to be
(1)
$$\begin{array}{cc}\hfill J_X(q,\mathrm{}):=& e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\underset{\beta }{}q^\beta J_X(\beta )\hfill \\ \hfill :=& e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\underset{\beta }{}q^\beta \mathrm{ev}_{}^{\mathrm{vir}}\frac{1}{\mathrm{}(\mathrm{}\psi )}\hfill \\ \hfill =& e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\underset{mk}{}T_mg^{mk}\underset{\beta }{}q^\beta _{[\overline{M}_{0,1}(X,\beta )]^{\mathrm{vir}}}\mathrm{ev}^{}(T_k)\frac{1}{\mathrm{}(\mathrm{}\psi )}\hfill \end{array}$$
and for $`EX`$:
(2)
$$\begin{array}{cc}\hfill J_X^E(q,\mathrm{}):=& e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\underset{\beta }{}q^\beta J_X^E(\beta ):=e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\mathrm{ev}_{}^{\mathrm{vir}}\frac{c_{\mathrm{top}}(E_\beta )}{\mathrm{}(\mathrm{}\psi )},\hfill \end{array}$$
which will be called *$`J`$-function* of $`X`$ and $`EX`$ respectively. It is easy to see that
$$_XJ_X^E=\underset{\beta }{}q^\beta _{[\overline{M}_{0,1}(X,\beta )]^{\mathrm{vir}}}\frac{c_{\mathrm{top}}(E_\beta )}{\mathrm{}(\mathrm{}\psi )}\mathrm{ev}^{}e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}.$$
For a line bundle $`L:=i^{}(๐ช_P(l))`$ on $`X`$ define
$$H_\beta ^L:=\underset{k=0}{\overset{c_1(L),\beta }{}}(c_1(L)+k\mathrm{})$$
for $`L`$ convex and
$$H_\beta ^L:=\underset{k=c_1(L),\beta +1}{\overset{1}{}}(c_1(L)+k\mathrm{})$$
for $`L`$ concave. Here a vector bundle $`EX`$ is called *convex* if for any stable map $`f:CX`$, $`H^1(C,\pi _{}f^{}(E))=0`$. $`E`$ is *concave* if $`H^0(C,\pi _{}f^{}(E))=0`$ for any $`f`$. Introduce another generating 0 of GromovโWitten invariants on $`X`$ (modified by $`H^L`$โs)
(3)
$$I_X^E(q,\mathrm{}):=e^{\frac{1}{\mathrm{}}(t_0+_ip_it_i)}\underset{\beta }{}q^\beta J_X(\beta )\underset{j}{}H_\beta ^{L_j}$$
###### Theorem 1.
Let $`X`$ be a smooth projective variety embedded in $`P=^{r_i}`$ (see *Notations.*) and $`E=_jL_jX`$ be the sum of line bundles $`L_j`$ which are the pull-backs of convex and concave line bundles on $`P`$ such that
$$c_1(T_X)\underset{L_j\text{convex}}{}c_1(L_j)+\underset{L_jconcave}{}c_1(L_j)$$
are non-negative. Then
$$_XJ_X^E(q,\mathrm{})_XI_X^E(q,\mathrm{}),$$
where $``$ means equivalence up to a *mirror transformation*, which is a special kind of change of variables described in ยง 4.3.
In fact, there are many situations when the above mirror transformations are unnecessary. The following theorem contains the main examples.
###### Theorem 2.
$`_XJ_X^E=_XI_X^E`$ if
1. $`E`$ is concave and $`\mathrm{rank}(E)2`$.
2. $`E`$ is convex and $`c_1(T_X)_{L_j\text{convex}}c_1(L_j)`$ is Fano of index $`2`$.
3. Direct sum of the previous two cases.
In the case when $`E:=(_jL_j)`$, all $`L_j`$ are *convex*, Theorem 1 has the following interpretation.
###### Corollary 1.
Let $`i_Y:YX`$ be the smooth zero locus of a section of $`E`$ (i.e. a complete intersection in $`X`$).
* 1.
$$_X(i_Y)_{}J_Y=_XJ_X^E.$$
In particular, Theorem 1 and Theorem 2 relate the GromovโWitten invariants of $`Y`$ to GromovโWitten invariants of $`X`$.
* 2. Suppose that $`H^{}(X)`$ is generated by divisor classes, and $`\mathrm{rk}(E)=1`$, i.e. $`Y`$ is a hypersurface. Then all $`n`$-point gravitational GromovโWitten invariants of $`i_Y^{}H^{}(X)`$ can be reconstructed from one-point gravitational GromovโWitten invariants of $`X`$.
###### Proof.
The part 1 is nothing but the statement that for any $`\omega H^{}(X)`$
(4)
$$[\overline{M}_{0,1}(Y,\beta )]^{\mathrm{vir}},i_Y^{}(\omega )=[\overline{M}_{0,1}(X,\beta )]^{\mathrm{vir}}c_{\mathrm{top}}(E_\beta ),\omega ,$$
which can be found in e.g. and references therein. See also ยง 1.2. The second part is a corollary of the first part plus a reconstruction theorem proved in , which states that one can reconstructs $`n`$-point descendants provided that $`H^{}(X)`$ is generated by divisor classes and one-point descendants are known. โ
###### Remark 1.
(Local mirror conjecture) When $`E`$ is concave, there is also an interpretation of the GromovโWitten invariants of $`EX`$: the invariants of the total space of vector bundle $`EX`$. For example $`๐ช(1)๐ช(1)^1`$, is the โtubular neighborhoodโ of $`^1`$ embedded in a CalabiโYau threefold with normal bundle $`๐ช_^1(1)๐ช_^1(1)`$.
###### Remark 2.
Whenever $`X`$ carries a group action by $`G`$, one could carry out the whole work to $`H_G^{}(X)`$ (instead of $`H^{}(X)`$) without any change. In this case, only localization of $`^{}`$-action on graph space is needed.
Combining the quantum differential equation and the above theorems, one can easily see the following interesting phenomenon. When $`X`$ is a toric variety this is a corollary of Giventalโs *quantum Serre duality* theorem .
###### Corollary 2.
Let $`E=_{j=1}^{\mathrm{rk}(E)}L_j`$ be a concave bundle with $`\mathrm{rk}(E)2`$. Then<sup>1</sup><sup>1</sup>1The restriction $`\mathrm{rk}(E)2`$ is shown unnecessary in the joint work with A. Bertram.
(5)
$$_Xc_{\mathrm{top}}(E)e^{\frac{1}{\mathrm{}}(t_0+pt)}J_X^E(q,\mathrm{})_X\frac{1}{c_{\mathrm{top}}(E^{})}e^{\frac{1}{\mathrm{}}(t_0+pt)}J_X^E^{}(q,\mathrm{}),$$
where $`E^{}`$ is the dual vector bundle and $``$ means equivalence up to mirror transformations (and a factor of power series in $`q`$).
An interesting consequence of (5) is that one can prove the mirror conjecture of convex bundles by concave bundles. For example the proof of mirror conjecture in the case of quintic three-fold can be carried out<sup>2</sup><sup>2</sup>2Here we are not very precise. (5) is valid only if $`\mathrm{rk}(E)2`$. However, we could use $`E=๐ช(5)๐ช(1)`$ on $`^5`$ instead of $`๐ช(5)`$ on $`^4`$. The reason is that quintic three-fold can be described as complete intersections in $`^5`$ of the bundle $`๐ช(1)๐ช(5)`$. by using $`๐ช(5)`$ instead of $`๐ช(5)`$ on $`^4`$. See ยง5 for an example. Here of course, this comes as a cyclic argument as we have used the proof of mirror conjecture in convex case to deduce this result. It is therefore desirable to have a direct proof of (5) and refine the statement. We plan to elaborate on this in a future paper (jointly with A. Bertram).
### 1.2. Relation to quantum cohomology
Lefschetz hyperplane theorem (LHT) asserts that a projective smooth variety $`X`$ contains essential (co-)homological information of its hyperplane section $`Y`$. The quantum Lefschetz hyperplane theorem (QLHT) verifies this assertion in quantum cohomology. It was proposed by A. Givental and formulated in the present form by B. Kim .
More precisely, our theorems state that the $`J`$-function of the complete intersection $`Y`$ (of classes $`i_Y^{}H^{}(X)`$) can be obtained from the $`J`$-function of $`X`$ by multiplying suitable cohomology classes and possibly a well-defined change of variables. The central roles of $`J`$-function in quantum cohomology theory is explained by the theory of *quantum differential equation* developed by Dijkgraaf and Givental . It says, first of all, that $`J`$-function is a flat section of the Dubrovin connection on the A-model side, parallel to the Picard-Fuchs equation on the B-model side. This is related to the mirror symmetry discussed in the next subsection. Secondly, it gives a nice way to obtain the essential information of small quantum ring $`QH^{}(X)`$ from $`J`$-function (*quantum $`๐`$-module*). For example, one can easily obtain the relations in quantum cohomology from $`J`$-function. Moreover, a result in states that it is possible to reconstruct n-point descendants from one-point descendants when $`H^{}(X)`$ is generated by divisor classes. Therefore our theorem even implicitly relates their big gravitational quantum cohomology algebras under this condition.
Many important special cases of QLHT, including the celebrated quintic three-fold, have been worked out by A. Givental ( and references therein), B. Kim and LiuโLianโYau . In the case of quintic three-fold, the ambient space $`X=^4`$ and $`E=๐ช(5)`$. Since it is very easy to compute the GromovโWitten invariants of $`^4`$, QLHT is therefore the central part of the proof of mirror conjecture which will be discussed in the next subsection.
### 1.3. Relation to mirror conjecture
In a seminal paper Candelas, de la Ossa, Green and Parkes applied the mirror symmetry to the quintic three-fold and derived, in a string-theoretic way, the celebrated formula which predicts the number $`n_d`$ of rational curves on the quintic three-fold of any degrees. This formula was then named *mirror conjecture*, or mirror identity to distinguish itself from more fundamental physical principle of mirror symmetry. This conjecture basically says that a generating function of $`n_d`$ is equivalent to a hypergeometric series up to a mirror transformation. Their result soon stimulated a lot of mathematical work in enumerative geometry. Among different groups working on the proof of their prediction, there have been notably two different approaches. One approach is trying to mathematically justify the string-theoretic mirror symmetry and therefore obtain mirror conjecture as a corollary. The other is to attack the enumerative consequence directly by developing new mathematics inspired from physics. GromovโWitten theory is partly inspired by this second approach which we will give a brief discussion<sup>3</sup><sup>3</sup>3The following is not meant to be a precise historical account..
The first major progress came from M. Kontsevich . As is well known in algebraic geometry, an enumerative problem can usually be formulated as an intersection-theoretic one on suitable moduli spaces. Kontsevich introduced the moduli space of stable maps and formulated the mirror conjecture as follows. Let $`๐ช(5)_d^{}:=\pi _{}\mathrm{ev}^{}(๐ช(5))`$ be the vector bundle on $`\overline{M}_{0,0}(^4,d)`$, then the enumerative problem was equivalent to computing the integral
(6)
$$N_d:=_{\overline{M}_{0,0}(^4,d)}c_{\mathrm{top}}(๐ช(5)_d^{}),$$
and $`N_d`$ can be related to $`n_d`$ by Aspinwall-Morrison Formula. By using torus action on $`^4`$ and fixed point localization method, he was able to reduce the integral (6) to summation of trees, but failed to complete the complicated combinatorial problem. Another (conceptual) drawback of this approach was that it did not explain the presence of hypergeometric series.
Then came A. Giventalโs proof followed by other approaches and generalizations by Lian-Liu-Yau and Bertram. The new innovations include, among other things, the introduction of equivariant quantum cohomology and graph space. The quantum cohomology of Calabi-Yau manifold $`X`$ is not semisimple, which makes the structure of quantum ring, like associativity relation, not very useful in computing $`n_d`$. By introducing equivariant quantum cohomology one produces a family of Frobenius structure over $`H_G^{}(pt)`$ whose generic fibre carries semisimple Frobenius structure while the special fibre $`H^{}(Y)`$ does not. Therefore $`QH^{}(Y)`$ for Calabi-Yau manifold $`Y`$ may be considered as a limiting case of semisimple Frobenius manifold. This explains, in one way, why the structure of quantum cohomology of Calabi-Yau manifolds play a role in enumerative problem. On the other hand, to properly explain the presence of hypergeometric series, the graph space $`\overline{G}_{0,0}(^r,d)`$ was introduced and was shown to have a natural birational morphism $`u`$ to the toric compactification space, or linear sigma model, $`_d^r:=^{(r+1)d+r}`$ (see ยง 2.1 for details). It was earlier found by E. Witten and Givental, etc. (from different approaches) that some suitable correlators on $`_d^r`$ actually produce the desired hypergeometric series. However, neither $`\overline{G}_{0,0}(^r,d)`$ nor $`_d^r`$ is the right space to perform the integral (6). The way to resolve this issue was to identify $`\overline{M}_{0,1}(^r,d)`$ as a fixed point component of $`^{}`$-action on $`\overline{G}_{0,0}(^r,d)`$. One then uses the birational morphism $`u:\overline{G}_{0,0}(^r,d)_d^r`$ to pass the above correlators from $`_d^r`$ to $`\overline{G}_{0,0}(^r,d)`$ and then pass to $`\overline{M}_{0,1}(^r,d)`$. In the quintic three-fold case, $`r=4`$ and the correlator obtained from this procedure is
(7)
$$\begin{array}{cc}\hfill J_^4^{๐ช(5)}(d):=& \mathrm{ev}_{}\frac{c_{\mathrm{top}}(๐ช(5)_d)}{\mathrm{}(\mathrm{}\psi )}\hfill \\ \hfill =& \mathrm{}^2\mathrm{ev}_{}c_{\mathrm{top}}(๐ช(5)_d)+\mathrm{}^3\mathrm{ev}_{}(c_{\mathrm{top}}(๐ช(5)_d)\psi )+\mathrm{}.\hfill \end{array}$$
The $`\mathrm{}^2`$ term in Laurent series expansion, when integrated over $`^4`$, will be (see )
$$_^4\mathrm{ev}_{}c_{\mathrm{top}}(๐ช(5)_d)=_{\overline{M}_{0,1}(^4,d)}c_{\mathrm{top}}(๐ช(5)_d)=_{\overline{M}_{0,0}(^4,d)}c_{\mathrm{top}}(๐ช(5)_d^{})$$
which is exactly (6).
There are now four approaches to mirror conjecture (known to us) by Givental, Lian-Liu-Yau, Bertram, and Gathmann. The interested reader can find valuable information in and references therein.
###### Remark 3.
Compare (7) with (1), one sees that our results can be interpreted as a generalization of mirror conjecture.
###### Acknowledgements.
My special thanks go to Aaron Bertram. The current proof is based on his work and I benefit a lot from our collaboration. I am also thankful to Bumsig Kim<sup>4</sup><sup>4</sup>4B. Kim informed us that he had previously obtained a special case of Theorem 2 case 2., Rahul Pandharipande for numerous useful discussions.
###### Update.
Gathmann has recently posted his proof of mirror theorem , where he proved the mirror theorem in the case $`E`$ is a convex line bundle (i.e. when $`Y`$ is a very ample hypersurface). The relation between his approach and ours is, roughly, the following. While we tried to sweep the classes $`e_\nu `$ in Theorem 5 โunder the carpetโ by dimensional constraints, he explicitly studies these relative classes and found a nice formula to relate these classes to ordinary GromovโWitten classes.
## 2. Graph space in GromovโWitten theory
*The Picard number of $`X`$ is assumed to be one throughout the rest of the paper, to avoid complicated notations. The generalization to the arbitrary Picard number is usually a matter of bookkeeping and is left to the reader.*
### 2.1. Graph space and one-point invariants
The $`n`$-pointed *graph space* of $`X`$ of degree $`\beta `$ is defined to be $`\overline{G}_{0,n}(X,\beta ):=\overline{M}_{0,n}(X\times ^1,(\beta ,1))`$, where the degree $`(\beta ,1)`$ is the element in $`H_2(X)H_2(^1)`$. It is a compactification of the space of maps from parameterized $`^1`$ to $`X`$. This space $`\overline{G}_{0,n}(X,\beta )`$ carries a natural $`^{}`$-action induced from the action on $`^1`$.
When $`X=^r`$ there is another (toric) compactification, $`_d^r`$ (*linear sigma model*), of the space of parameterized map of degree $`d`$. It is constructed in the following way. Consider the projective space of $`(r+1)`$-tuple of the degree $`d`$ (symmetric) binary forms of $`(z_0:z_1)`$. When there is no common factors of positive degrees among these $`(r+1)`$-tuples, it represents a morphism from $`^1^r`$. We may compactify it by simply allowing the common factors and taking quotient by $`^{}`$-action. It is easy to see that this space is equal to $`^{(r+1)(d+1)1}`$.
By construction $`\overline{G}_{0,0}(^r,d)`$ is birationally isomorphic to $`_d^r`$. In fact,
###### Theorem 3.
*(Giventalโs Main Lemma )* There exists a natural birational $`^{}`$-equivariant morphism $`u:\overline{G}_{0,0}(^r,d)_d^r`$.
This morphism $`u`$ can be described as follows. Consider a stable degree $`(d,1)`$ map $`f:C^1\times ^n`$. There exists a unique irreducible component $`C_0C`$ (called *parameterized component*) such that $`f|_{C_0}`$ has degree $`(d_0,1)`$ where $`d_0d`$. The image $`f(C_0)`$ is the graph of a map $`^1^r`$ of degree $`d_0`$. The map is given by the binary forms $`(p_0:\mathrm{}:p_r)`$ of degree $`d_0`$ with no common factors and determines the forms uniquely up to a non-zero constant factor. The curve $`\overline{CC_0}`$ has $`s`$ connected *unparameterized components* which are mapped to $`^1\times ^r`$ with degrees $`(d_1,0),\mathrm{},(d_s,0),d_1+\mathrm{}+d_s=dd_0`$, and the image of $`i`$-th component is contained in the slice $`(a_i:b_i)\times ^n`$. We put $`u(C,f)=_{i=1}^r(a_iz_0b_iz_1)^{d_i}(p_0:\mathrm{}:p_r)`$. For a detailed proof in algebro-geometric terms see and .
Of course we can also find simple birational models for $`\overline{G}_{0,s}(^r,d)`$. A particular useful one is $`(^1)^s\times _d^r`$. There is also a morphism $`u_s:\overline{G}_{0,s}(^r,d)(^1)^s\times _d^r`$. The first factor is defined by the composition
$$\overline{G}_{0,s}(^r,d)\stackrel{\mathrm{ev}_i}{}(X\times ^1)^s\stackrel{p_2}{}(^1)^s$$
and the second factor is the composition of the forgetful morphism $`\overline{G}_{0,s}(^r,d)\overline{G}_{0,0}(^r,d)`$ and $`u:\overline{G}_{0,0}(^r,d)_d^r`$ defined in Theorem 3.
From the above description, it is clear that two spaces $`\overline{G}_{0,0}(^r,d)`$ and $`_d^r`$ differ on certain boundary strata. The *comb type strata*, denoted $`D_\mu `$ where $`\mu :=(d_0,d_1,\mathrm{},d_s),d_0+d_1+\mathrm{}d_s=d`$, play an important role in our discussion. The (domain) curve of a generic element in $`D_\mu `$ has one parameterized component and $`s`$ nodes, and $`(d_0,1)H_2(^r\times ^1)`$ is the degree of the parameterized component. These strata have substrata, called *hairy comb type strata*, which are obtained by further degenerating the unparameterized components. Note that different permutations of $`d_1,\mathrm{},d_s`$ represent the same $`\mu `$. For example $`(d_0,d^{},d^{\prime \prime })=(d_0,d^{\prime \prime },d^{})`$. For such a stratum $`D_\mu `$ there is a finite birational morphism
$$\stackrel{~}{D}_\mu :=\overline{G}_{0,s}(^r,d_0)\times _{(^r)^s}\underset{m=1}{\overset{s}{}}\overline{M}_{0,1}(^r,d_m)D_\mu .$$
To simplify our notation, we will denote $`\stackrel{~}{D}_\mu `$ also by $`D_\mu `$ henceforth. There are also useful morphisms from $`D_\mu `$ to simple spaces:
(8)
$$u_\mu :D_\mu =\overline{G}_{0,s}(^r,d_0)\times _{(^r)^s}\underset{m=1}{\overset{s}{}}\overline{M}_{0,1}(^r,d_m)(^1)^s\times _{d_0}^r,$$
defined by the composition
(9)
$$D_\mu \stackrel{p_0}{}\overline{G}_{0,s}(^r,d_0)\stackrel{u_s}{}(^1)^s\times _{d_0}^r.$$
Let $`X\stackrel{๐}{}^r`$ be an embedding described earlier. For the notational convenience, we define the divisors $`D_\nu :=i_G^{}D_\mu `$ of $`\overline{G}_{0,0}(X,\beta )`$ (indexed by $`\nu :=(\beta _0,\beta _1,\mathrm{},\beta _s)`$) for future reference. The relation is described in the following commutative diagram:
(10)
The reason to introduce the graph space in GromovโWitten theory is to obtain one-point descendant invariants. The graph space $`\overline{G}_{0,0}(X,\beta )`$ carries a $`^{}`$ action induced from the $`^{}`$ action on $`^1`$, and $`\overline{M}_{0,1}(X,\beta )`$ is a fixed point component. It is summarized in the following commutative diagram:
(11)
where the left upper square is a commutative diagram such that $`\overline{M}_{0,1}(X,\beta )\overline{G}_{0,0}(X,\beta )`$ and $`\overline{M}_{0,1}(P,d)\overline{G}_{0,0}(P,d)`$ are fixed point components of $`^{}`$-action on the graph spaces and $`\overline{M}_{0,1}(X,\beta )`$ is the only fixed point component mapping to $`\overline{M}_{0,1}(^r,d)`$ by $`i_M`$. The same can be said about the right square: $`^r_d^r`$ is a fixed point component such that $`\overline{M}_{0,1}(^r,d)`$ is the only fixed point component mapping to $`^r`$.
###### Remark 4.
The above setting actually works for product of projective spaces $`P:=^{r_i}`$. There are only small changes in this adjustment. First, all $`d,\mu `$ etc. should stand for multi-index. Second, the birational morphism in Theorem 3 should be replaced by
$$u:\overline{G}_{0,0}(P,d)\underset{i}{}_{d^i}^{r_i}.$$
### 2.2. Virtual localization on graph space
First recall the Graber-Pandharipande virtual localization formula :
###### Theorem 4.
Let $`X`$ be an algebraic scheme with a $`^{}`$ action and $`^{}`$-equivariant perfect obstruction theory. Then the virtual localization formula holds:
$$[X]^{\mathrm{vir}}=j_{}\frac{[X_j]^{\mathrm{vir}}}{e(N_{X_j|X}^{\mathrm{vir}})}$$
in $`A_{}^{^{}}(X)[\lambda ,\frac{1}{\lambda }]`$, where $`\lambda `$ is the generator of the $`A_{}^{^{}}(pt)`$.
An immediate consequence of this theorem is the *correspondence of residues* ( Lemma 2.1 and ):
###### Corollary 3.
Suppose that $`f:X_1X_2`$ is a $`^{}`$-equivariant map of two algebraic schemes and $`j_1:F_1X_1`$ and $`j_2:F_2X_2`$ are two fixed point components of $`X_1`$ and $`X_2`$ respectively, such that $`F_1`$ is the only fixed point component mapping into $`F_2`$ as in the following commutative diagram:
$$\begin{array}{ccc}X_1& \stackrel{f}{}& X_2\\ j_1& & j_2& & \\ F_1& \stackrel{f|_{F_1}}{}& F_2.\end{array}$$
Then
(12)
$$f|_{F_1}^{}{}_{}{}^{}\left(\frac{j_1^{}(\omega )[F_1]^{\mathrm{vir}}}{e(N_{F_1|X_1}^{\mathrm{vir}})}\right)=\frac{j_2^!f_{}(\omega [X_1]^{\mathrm{vir}})}{e(N_{F_2|X_2}^{\mathrm{vir}})}$$
for any $`\omega H_{^{}}^{}(X_1)`$.
Apply this result to our case: Let $`X_1=\overline{G}_{0,0}(X,\beta )`$, $`X_2=\overline{G}_{0,0}(P,d)`$, $`F_1=\overline{M}_{0,1}(X,\beta )`$, $`F_2=\overline{M}_{0,1}(P,d)`$ and $`\omega =c_{\mathrm{top}}(E_\beta ^G)`$, as displayed in the upper left square of (11). Since $`j_X^{}c_{\mathrm{top}}(E_\beta ^G)=c_{\mathrm{top}}(E_\beta )`$, one has
(13)
$$\begin{array}{cc}\hfill i_{M}^{\mathrm{vir}}{}_{}{}^{}\left(\frac{c_{\mathrm{top}}(E_\beta )}{\mathrm{}(\mathrm{}\psi )}\right)=& \mathrm{PD}i_{M}^{}{}_{}{}^{}\left(\frac{c_{\mathrm{top}}(E_\beta )}{\mathrm{}(\mathrm{}\psi )}[\overline{M}_{0,1}(X,\beta )]^{\mathrm{vir}}\right)\hfill \\ \hfill =& \frac{\mathrm{PD}j_P^!i_{G}^{}{}_{}{}^{}\left(c_{\mathrm{top}}(E_\beta ^G)[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}\right)}{\mathrm{}(\mathrm{}\psi )}\hfill \\ \hfill =& \frac{j_P^{}i_{G}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}(E_\beta ^G)}{\mathrm{}(\mathrm{}\psi )},\hfill \end{array}$$
where $`\mathrm{}`$ is the generator of $`H_{^{}}^{}(pt)`$ and $`\psi `$ is the first chern class of the tautological line bundle $`_1`$ on $`\overline{M}_{0,1}(X,\beta )`$ or $`\overline{M}_{0,1}(P,d)`$. The one small difference between (12) and (13) is that $`\overline{G}_{0,0}(P,d)`$ and $`\overline{M}_{0,1}(P,d)`$ are orbifolds and Poincarรฉ duality makes sense there.
###### Remark 5.
The functorial properties of virtual fundamental classes used in this article can be found in .
## 3. Decomposition of the virtual fundamental classes
Recall that there is a birational morphism $`u_{X,1}:=u_1i_G`$ (see (8)) from the universal curve $`\overline{G}_{0,1}(X,\beta )`$ of the graph space $`\overline{G}_{0,0}(X,\beta )`$ to $`^1\times _d^r`$. Given a line bundle $`L:=i^{}๐ช_^r(l)`$ on $`X`$ one could produce two line bundles $`\mathrm{ev}^{}L`$ and $`u_{X,1}^{}๐ช_{^1\times _d^r}(dl,l)`$ on $`\overline{G}_{0,1}(X,\beta )`$. These two line bundles are isomorphic on the open subset $`U:=\overline{G}_{0,1}(X,\beta )_{\nu =(\beta _0,\beta _1)}C_\nu `$, where $`C_\nu `$ is the universal curve of the *unparameterized* component over $`D_\nu `$. The reason is that $`C_\nu `$ are exactly the exceptional divisors of $`u_1`$. More explicitly, there is a rational map $`b:^1\times _d^r^r`$ such that $`b`$ is well-defined on $`u(U)`$ in the following commutative diagram:
It is easy to see that $`b`$ is a morphism on $`u(U)`$, of degree $`d`$ in the first factor and linear in the second. Namely
$$b|_{u(U)}^{}๐ช_^r(l)=๐ช_{^1\times _d^r}(dl,l).$$
Therefore
$$\mathrm{ev}_X^{}(L)=u_{X,1}^{}(๐ช_{^1\times _d^r}(dl,l))๐ช\left(\underset{\nu =(\beta _0,\beta _1)}{}c_\nu C_\nu \right).$$
It is easy to see that $`c_\nu `$ is negative as $`\mathrm{ev}_X^{}(L)๐ช_{^1\times _d^r}(dl,l))`$ by restricting a section of $`\mathrm{ev}_X^{}(L)`$ to $`U`$ and then extending this section on $`_d^r`$ by Hartogโs lemma. The actual coefficients<sup>5</sup><sup>5</sup>5In fact, we will only need to know $`c_\nu `$ is negative. $`c_\nu `$ can be determined by the following observation. The degree of $`f^{}(L)`$ on the *unparameterized* component of the universal curve over the divisor $`D_\nu `$ (at a generic point) has degree $`c_1(L),\beta _1`$. This leads to $`c_\nu =c_1(L),\beta _1`$. The same argument applies to $`n`$-pointed graph space:
###### Lemma 1.
On the universal curve $`C`$ over the graph space $`\overline{G}_{0,n}(X,\beta )`$
$$\begin{array}{cc}& \mathrm{ev}_{1,X}^{}(L)\hfill \\ \hfill =& u_{X,n}^{}(๐ช_{^1\times (^1)^n\times _d^r}(dl,0,\mathrm{},0,l))๐ช\left(\underset{\nu =(\beta _0,\beta _1)}{}c_1(L),\beta _1C_\nu \right),\hfill \end{array}$$
where $`C_\nu `$ is the universal curve of the unparameterized component over $`D_\nu `$.
###### Corollary 4.
For $`L`$ convex let
$$L_\beta ^G:=R^0\pi _{n+1}\mathrm{ev}_{n+1}^{}(L_l)$$
$$F_\beta ^G:=u_X^{}\left(H^0(^1,๐ช(dl))๐ช_{(^1)^n\times _d^r}(0,\mathrm{},0,l)\right).$$
One has equivariant maps of vector bundles on graph space $`\overline{G}_{0,n}(X,\beta )`$
$$\sigma _0:L_\beta ^GF_\beta ^G.$$
For $`L`$ concave let
(14a)
$$L_\beta ^G:=R^1\pi _{}e^{}(L_l(x_1\mathrm{}x_n))$$
(14b)
$$F_\beta ^G:=u_X^{}\left(H^1(^1,๐ช(dl)(\chi _1\mathrm{}\chi _n))๐ช_{(^1)^n\times _d^r}(0,l)\right).$$
One has
$$\sigma _1:F_\beta ^GL_\beta ^G.$$
Here in (14b) $`\chi _m:(^1)^n_d^r^1`$ are the projections to the $`m`$-th factor of $`^1`$ (considered as โmarked pointsโ). In equation (14a), $`x_m`$ are the marked points on the universal curve.
###### Proof.
It is clear that the vector bundles in this lemma are the push-forwards of two line bundles on the universal curve $`C`$ considered in Lemma 1. When the coefficient $`c_\nu `$ is negative, one has the following inclusion
(15)
$$\mathrm{ev}_X^{}(L)u_{X,n}^{}๐ช_{^1\times (^1)^n\times _d^r}(dl,0,d).$$
$`\sigma _0`$ is then obtained by pushing-forward the above inclusion of line bundles to $`\overline{G}_{0,n}(X,\beta )`$.
$`\sigma _1`$ can be obtained in a similar way. When $`L`$ is concave, $`c_1(L),\beta _1`$ is positive so that the arrow of (15) is reversed. โ
The equation (4) indicates that the QLHT boils down to the study of the top chern class $`c_{\mathrm{top}}(E_\beta )`$ on $`\overline{M}_{0,1}(X,\beta )`$, which is in turn the pull-back $`j_X^{}(c_{\mathrm{top}}(E_\beta ^G))`$ of top chern class from $`\overline{G}_{0,0}(X,\beta )`$. From the above discussion
$$c_{\mathrm{top}}(E_\beta ^G)=u_X^{}c_{\mathrm{top}}(F_\beta ^G)+\text{boundary terms},$$
where the boundary terms are supported on the comb type strata $`D_\nu `$. Notice that the โmain termโ $`c_{\mathrm{top}}(F_\beta ^G)`$ is the (equivariant) top chern class of direct sum of line bundles on the projective space and can therefore be easily computed. In fact, we will see in the next section that this part gives rise to the factors $`H_\beta ^E`$ in (3). It remains to have a close look of the boundary terms. A key observation of is that boundary contributions can be explicitly computed using MacPhersonโs graph construction for vector bundle morphisms.
###### Proposition 1.
When $`X=^r`$ and $`E=๐ช_^r(l)`$ a line bundle,
$$c_{\mathrm{top}}(๐ช_^r(l)_d^G)=u^{}\underset{k=0}{\overset{dl}{}}(lH+k\mathrm{})+\underset{\mu }{}\frac{1}{s!}\phi _{\mu }^{}{}_{}{}^{}\left(e_\mu u_\mu ^{}\underset{k=0}{\overset{d_0l}{}}(lH+k\mathrm{})\right),$$
for $`l`$ *positive* and
$$c_{\mathrm{top}}(๐ช_^r(l)_d^G)=u^{}\underset{k=dl+1}{\overset{1}{}}(lH+k\mathrm{})+\underset{\mu }{}\frac{1}{s!}\phi _{\mu }^{}{}_{}{}^{}\left(e_\mu u_\mu ^{}\xi _1\mathrm{}\xi _s\underset{k=d_0l+1}{\overset{1}{}}(lH+k\mathrm{})\right),$$
for $`l`$ negative. Here $`s`$ is the number of unparameterized components for a generic curve over $`D_\nu `$, $`H`$ is the hyperplane class of $`_d^r`$ and $`\xi _m`$ is the point class of $`m`$-th $`^1`$ in $`(^1)^s\times _d^r`$. The equivariant class $`e_\mu `$ is defined in (23).
###### Theorem 5.
Let $`E=_jL_j`$, $`L_j=i^{}(๐ช_^r(l_j))`$. The virtual fundamental class $`c_{\mathrm{top}}(E_\beta ^G)[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}`$ decomposes as follows:
For *convex* bundle $`E`$
(16)
$$\begin{array}{cc}& c_{\mathrm{top}}(E_\beta ^G)[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}\hfill \\ \hfill =& u_X^{}\underset{j}{}\underset{k_j=0}{\overset{c_1(L_j),\beta }{}}(l_jH+k_j\mathrm{})[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}\hfill \\ \hfill +& \underset{\nu }{}\frac{1}{s!}\phi _{\nu }^{}{}_{}{}^{}\left(\left(e_\nu u_\nu ^{}\underset{j}{}\underset{k_j=0}{\overset{c_1(L_j),\beta _0}{}}(l_jH+k_j\mathrm{})\right)[D_\nu ]^{\mathrm{vir}}\right).\hfill \end{array}$$
For *concave* $`E`$
$$\begin{array}{cc}& c_{\mathrm{top}}(E_\beta ^G)[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}\hfill \\ \hfill =& u_X^{}\underset{j}{}\underset{k_j=c_1(L_j),\beta +1}{\overset{1}{}}(l_jH+k_j\mathrm{})[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}\hfill \\ \hfill +& \underset{\nu }{}\frac{1}{s!}\phi _{\nu }^{}{}_{}{}^{}\left(\left(e_\nu u_\nu ^{}\underset{j}{}\underset{k_j=c_1(L_j),\beta _0+1}{\overset{1}{}}\xi _1\mathrm{}\xi _s(l_jH+k_j\mathrm{})\right)[D_\nu ]^{\mathrm{vir}}\right).\hfill \end{array}$$
In the case $`\mathrm{rk}(E)=1`$, $`e_\nu :=i_G^{}e_\mu `$. In general, it is defined inductively.
###### Proof.
It is easy to see that the theorem follows from the above proposition by the formula $`c_{\mathrm{top}}(E_\beta ^G)=_jc_{\mathrm{top}}((L_j)_\beta ^G)`$ and the excess intersection theory. Note that $`c_{\mathrm{top}}((L_j)_\beta ^G)=i_G^{}c_{\mathrm{top}}(๐ช_^r(l_j)_d^G)`$. โ
###### Proof.
(of Proposition 1)
I. (*Convex case*) The convex case of the proposition is Lemma 4.4 of . For the convenience of the reader and future references, we reproduce Bertramโs proof.
As remarked above, it is clear from Lemma 1 that the equivariant virtual class has its โmainโ contribution from
$$c_{\mathrm{top}}(u^{}๐ช_{^1\times _d^r}(ld,l))=\left(u^{}\underset{k=0}{\overset{ld}{}}(lH+k\mathrm{})\right).$$
It is also clear that other contributions to $`c_{\mathrm{top}}(๐ช_^r(l)_d^G)`$ come from the boundary strata $`\mu `$ (and possibly its substrata). It remains to study the boundary terms. This can be done by MacPhersonโs graph construction of vector bundle morphisms.
Let $`_d:=๐ช_^r(l)_d^G`$ and $`_d:=H^0(^1,๐ช(dl))u^{}๐ช_{_d^r}(l)`$ be vector bundles on $`\overline{G}_{0,0}(^r,d)`$. By Corollary 4 there is a homomorphism $`\sigma _0:_d_d`$ of vector bundles of the same rank $`dl+1`$. Let $`G:=\text{Grass}_{dl+1}(_d_d)`$ be the Grassmann bundle over $`\overline{G}_{0,0}(^r,d)`$, with universal bundle $`\zeta G`$ of rank $`dl+1`$. There is a canonical embedding
$$\mathrm{\Phi }:\overline{G}_{0,0}(^r,d)\times ๐ธ^1G\times ^1$$
taking $`(z,\lambda )`$ to $`(\text{graph of }\lambda \sigma _0(z),(1:\lambda ))`$. Let $`W`$ be the closure of the image of $`\mathrm{\Phi }`$ and let
$$W_{\mathrm{}}=i_{\mathrm{}}^{}[W]=m_\delta [V_\delta ]$$
be a cycle in $`G`$ of dimension equal to $`dim(\overline{G}_{0,0}(^r,d))`$, where $`m_\delta `$ is the multiplicity of $`V_\delta `$. Let
$$\eta _\delta :V_\delta GZ_\delta \overline{G}_{0,0}(^r,d)$$
be the map induced by projection, and $`Z_\delta `$ be the image of $`\eta _\delta `$. The philosophy of the graph construction is that different components $`Z_\delta `$ are responsible for different types of degeneration of $`\sigma _0`$. Since $`\sigma _0`$ is generically of full rank, there is one component $`V_0Z_0\overline{G}_{0,0}(^r,d)`$ in $`W_{\mathrm{}}`$ with multiplicity one such that $`Z_0`$ embeds in $`G`$ via the fibre of $`_d`$. Other $`V_\delta `$ map to proper sub-varieties $`Z_\delta `$ of $`\overline{G}_{0,0}(^r,d)`$. It follows that (, Example 18.1.6)
(17)
$$c_{\mathrm{top}}(_d)=u^{}\underset{k=0}{\overset{dl+1}{}}(lH+k\mathrm{})+\underset{\delta }{}m_\delta \eta _{\delta }^{}{}_{}{}^{}c_{\mathrm{top}}(\zeta ).$$
To calculate the contributions from $`V_\delta `$, further study on the behavior of $`\sigma _0`$ on boundary strata $`D_\mu `$ is needed. Let $`f:C^r`$ be a generic stable map in strata $`\mu =(d_0,\mathrm{},d_r)`$. $`\sigma _0|_{D_\mu }`$ can be described fibrewisely (at a generic point) as:
(18)
$$H^0(C,f^{}๐ช_^r(l))H^0(C_0,f^{}๐ช_^r(l)|_{C_0})H^0(^1,๐ช(d_0l))H^0(^1,๐ช(dl)),$$
where $`C_0^1`$ is the parameterized component of $`C`$. The first map is simply the restriction and the last map is defined by multiplying a factor $`_{m=1}^s(a_mz_0b_mz_1)^{ld_m}`$, where $`(a_m:b_m)`$ are the nodal points on the parameterized $`^1`$ and $`(z_0:z_1)`$ the homogeneous coordinates on parameterized $`^1`$. It follows that $`\sigma _0`$ drops ranks only on the comb strata of types $`\mu `$ described in the previous section. Any hairy comb substrata obtained from $`\mu `$ by further degenerating the unparameterized components will not further reduce the rank of $`\sigma _0`$. Moreover, it has the following *transversality* property: $`\sigma _0`$ has generic corank $`n_1`$ and $`n_2`$ along $`D_{\mu _1}`$ and $`D_{\mu _2}`$ respectively. Then $`\sigma _0`$ has generic corank $`n_1+n_2`$ along the intersection of two strata.
The above fibrewise description actually holds globally. Namely, on the strata $`D_\mu `$, $`\sigma _0`$ is the following composition
(19)
$$\phi _\mu ^{}\sigma _0:\phi _\mu ^{}_dp_0^{}_{d_0}\stackrel{p_0^{}\sigma _0}{}p_0^{}_{d_0}\phi _\mu ^{}_d,$$
where $`p_0:D_\mu \overline{G}_{0,s}(^r,d_0)`$ is the projection (see (9)), $`_{d_0}`$ and $`_{d_0}`$ on $`\overline{G}_{0,s}(^r,d_0)`$ are defined in Corollary 4. The last map in (19) is the push-forward (along the first $`^1`$) of the following map
$$๐ช_{^1\times (^1)^s\times _{d_0}^r}(d_0l,0,\mathrm{},0,l)๐ช_{^1\times (^1)^s\times _{d_0}^r}(dl,d_1l,\mathrm{},d_sl,l).$$
One can now apply the above study of degeneration type of $`\sigma _0`$ to the graph construction. Since the rank of $`\sigma _0`$ decreases only on the strata $`D_\mu `$, (17) becomes
(20)
$$c_{\mathrm{top}}(_d)=u^{}\underset{k=0}{\overset{dl}{}}(lH+k\mathrm{})+\underset{\mu }{}m_\mu S_\mu .$$
Namely, the only $`Z_\delta `$ in (17) are $`D_\mu `$ and $`S_\mu =\eta _{\mu }^{}{}_{}{}^{}c_{\mathrm{top}}(\zeta )`$. To write down $`S_\mu `$ explicitly in terms of characteristic classes in $`\sigma _0|_{D_\mu }:\phi _\mu ^{}_d\phi _\mu ^{}_d`$ one would need a filtration of the above setting. Let us first deal with the simplest case when $`\mu =(d_0,d_1)`$, i.e. $`D_\mu `$ is a divisor. The universal curve on $`D_\mu =\overline{G}_{0,1}(^r,d_0)\times _^r\overline{M}_{0,1}(^r,d_1)`$ (generically) consists of one parameterized and one unparameterized components. The kernel of $`\sigma _0`$ can be identified with $`p_1^{}_{d_1}^1`$, where $`p_1:D_\mu \overline{M}_{0,1}(^r,d_1)`$ and $`_{d_1}^1`$ is the kernel of the evaluation morphism $`e`$ of bundles on $`\overline{M}_{0,1}(^r,d_1)`$
(21)
$$0_{d_1}^1\pi _2\mathrm{ev}_2^{}๐ช_^r(l)\stackrel{๐}{}\mathrm{ev}_1^{}๐ช_^r(l)0,$$
where $`\mathrm{ev}_2:\overline{M}_{0,2}(^r,d_1)^r`$ and $`\pi _2:\overline{M}_{0,2}(^r,d_1)\overline{M}_{0,1}(^r,d_1)`$ forgets the second marked point. The kernel $`_{d_1}^1`$ can be further filtered by the order of zeros of $`e`$:
(22)
$$0=p_1^{}_{d_1}^{d_1l+1}p_1^{}_{d_1}^{d_1l}\mathrm{}p_1^{}_{d_1}^1\phi _\mu ^{}_d$$
where $`_{d_1}^k`$ consists of those sections of $`\mathrm{ev}_2^{}๐ช_^r(l)`$ which vanishes at least to the $`k`$-th order at the marking. Similarly we can filter $`\phi _\mu ^{}_d`$ by the span of the image of $`\sigma _0|_{kD_\mu }`$ on $`^1\times _{d_0}^r`$:
$$_{d_0}^k:=H^0(^1,๐ช(d_0l+k1))๐ช_{^1\times _{d_0}^r}(k1,l)$$
such that
$$p_1^{}_{d_0}=u_\mu ^{}_{d_0}^1\mathrm{}u_\mu ^{}_\mu ^{d_1l}\phi _\mu ^{}_d$$
is a filtration of $`\phi _\mu ^{}_d`$ on $`D_\mu `$. Now for each $`D_\mu `$ there are $`d_1l`$ components $`V_\mu ^kW_{\mathrm{}}`$ ($`\eta _\mu :V_\mu D_\mu `$) because the infinitesimal property of $`\sigma _0`$ along $`D_\mu `$. Each of $`V_\mu ^k`$ is a birational image of a $`^1`$-bundle over $`D_\mu `$. By a local computation (), $`V_\mu ^k`$ has multiplicity<sup>6</sup><sup>6</sup>6The explicit multiplicity is actually irrelevant to our result. $`k`$, and the tautological bundle on $`V_\mu ^k`$ can be expressed in terms of the filtration of $`_d`$ and $`_d`$:
$$0p_0^{}u_\mu ^{}_{d_0}^kp_1^{}_{d_1}^{k+1}\zeta _{V_\mu ^k}๐ช(1)0.$$
The contribution $`S_\mu `$ from the boundary strata $`\mu =(d_0,d_1)`$ can be therefore written as
$$S_\mu =\phi _{\mu }^{}{}_{}{}^{}\left(e_\mu c_{\mathrm{top}}(u_\mu ^{}_{d_0}^1)\right)=\phi _{\mu }^{}{}_{}{}^{}\left(e_\mu u_\mu ^{}\underset{k=0}{\overset{d_0l}{}}(lH+k\mathrm{})\right)$$
where
$$e_\mu =\underset{k=1}{\overset{d_1l}{}}(k)p_{d_1}^{}c_{\mathrm{top}}(_{d_0}^{k+1})u_\mu ^{}c_{\mathrm{top}}(_{d_0}^k/_{d_0}^1).$$
This is exactly what we are looking for.
When $`\mu `$ is not a divisorial stratum we may use the above transversality property (of $`\sigma _0`$ concerning the intersection of two boundary strata). Since every $`D_\mu `$ is the intersection of divisors, and the corank of $`\sigma _0`$ at intersection is equal to the sum of the coranks generically, we can then obtain $`S_\mu `$ for general $`\mu `$ inductively. Let $`\mu _1=(d_0,d_1)`$ and $`V_{\mu _1}^kW_{\mathrm{}}`$ be the $`^1`$-bundle over $`D_{\mu _1}`$. Now apply graph construction to the following vector bundle morphism on (pulled back to) $`V_{\mu _1}^k`$
$$\phi _\mu ^{}_d/p_1^{}_{d_1}^1p_0^{}_{d_0}\stackrel{p_0^{}\sigma _0}{}p_0^{}_{d_0}=u_\mu ^{}_{d_0}^1.$$
(Some obvious pull-backs will be omitted.) Then the components of $`W_{\mathrm{}}^{}`$ over $`V_{\mu _1}^k`$ obtained from this construction map birationally to the components of $`W_{\mathrm{}}`$ over $`\overline{G}_{0,0}(^r,d)`$. For example, components of $`V_\mu ^{}^mW_{\mathrm{}}^{}`$ corresponding to boundary strata $`\mu ^{}=(d_0d^{},d^{})`$ of $`\overline{G}_{0,1}(^r,d_0)`$ map to the components of $`V_\mu ^mW_{\mathrm{}}`$ over the boundary strata $`\mu =(d_0d^{},d^{},d_1)`$ on $`\overline{G}_{0,0}(^r,d)`$. This implies that $`V_\mu `$ for any $`\mu =(d_0,d_1,\mathrm{},d_s)`$ with a fixed ordering of $`d_1,\mathrm{},d_s`$, $`V_\mu ^k`$ are (birationally) towers of $`^1`$-bundles over $`D_\mu `$ (by first doing $`\mu _1=(d_0+\mathrm{}+d_{s1},d_s)`$ then $`\mu _2=(d_0+\mathrm{}+d_{s2},d_{s1},d_s)`$, etc.). An explicit expression of $`e_\mu `$ can therefore be obtained:
(23)
$$e_\mu =\underset{m=1}{\overset{s}{}}\underset{k_m=1}{\overset{d_ml}{}}(k_m)p_{d_m}^{}c_{\mathrm{top}}(_{d_m}^{k_m+1})u_\mu ^{}c_{\mathrm{top}}(_{\mathrm{\Sigma }_m}^{k_m}/_{\mathrm{\Sigma }_m}^1)$$
where $`\mathrm{\Sigma }_m:=_{a=0}^{m1}d_a`$ and $`_{d_m}^k`$ is defined to be the filtration of kernel on the $`m`$-th unparameterized component, similar to that defined in (22) and $`_{\mathrm{\Sigma }_m}^1_{\mathrm{\Sigma }_m}^{k_m}`$ is the push-forward of the inclusion of line bundles on $`^1\times (^1)^s\times P_{d_0}`$ to $`(^1)^s\times P_{d_0}`$:
$$\begin{array}{cc}\hfill ๐ช& (\underset{a=1}{\overset{m1}{}}d_al,d_1l,\mathrm{},d_{m1}l,0,\mathrm{},0,l)\hfill \\ \hfill ๐ช& (\underset{a=1}{\overset{m1}{}}d_al+k_m1,d_1l,\mathrm{},d_{i1}l,k_m1,0,\mathrm{},0,l).\hfill \end{array}$$
This completes our proof of the convex case.
II. (*Concave case*) The proof of the concave case can in general be carried out in a similar way. However, some crucial modifications will be necessary. Now $`_d=R^1\pi _{}\mathrm{ev}^{}๐ช_^r(l)`$ and $`_d=H^1(^1,๐ช(dl))u^{}๐ช_{_d^r}(l)`$ (with $`l`$ negative). Corollary 3 guarantees that there is a bundle map $`\sigma _1`$ between $`_d`$ and $`_d`$. We may apply Serre duality and find $`\sigma _1^{}:(_d)^{}(_d)^{}`$. Carry out the graph construction to $`\sigma _1^{}`$. The equation (18) (on $`D_{(d_0,\mathrm{},d_s)}`$) should be replaced by
$$\begin{array}{c}H^0(C,\omega _Cf^{}L^1)H^0(C_0,\omega _{C_0}(\chi _1+\mathrm{}+\chi _s)f^{}L^1|_{C_0})\hfill \\ \hfill H^0(^1,\omega _^1(d_0l)๐ช(\chi _1+\mathrm{}+\chi _s))H^0(^1,\omega _^1(dl)).\end{array}$$
Here $`\chi _1,\mathrm{},\chi _s`$ are the nodal points on the parameterized $`^1`$ (see Corollary 3) and $`\omega _C`$ is the dualizing sheaf of $`C`$.
A similar modification to (20) should also take place:
$$c_{\mathrm{top}}(E_d)=u_X^{}\underset{k=c_1(L),d+1}{\overset{1}{}}(lH+k\mathrm{})+\underset{\mu }{}m_\mu S_\mu .$$
For the filtration of the kernel of $`\sigma _1^{}`$, we may use the following exact sequence (see (21))
$$0_{d_1}^1\pi _{}(\mathrm{ev}_2^{}๐ช_^r(l)\omega (x_1))\stackrel{res}{}\mathrm{ev}_1^{}๐ช_^r(l)0,$$
where $`\mathrm{res}`$ is the residue at $`x_1`$. Again we can further filter $`_{d_1}^1`$ by the order of zeros of $`\mathrm{res}`$, (as did in (22)). Similar filtration can be defined on $`\phi _\mu ^{}_d`$. Namely
$$\begin{array}{cc}\hfill _{d_0}^{k_m}=& \omega _^1\left(x_1+\mathrm{}+x_s+\underset{a=1}{\overset{m1}{}}d_a(l)+k_m1\right)\hfill \\ & ๐ช_{(^1)^s\times P_{d_0}}(d_1(l),\mathrm{},d_{m1}(l),k_m1,0,\mathrm{},0,(l)).\hfill \end{array}$$
Now a similar computation leads to
(24)
$$\begin{array}{cc}\hfill S_\mu =& \phi _{\mu }^{}{}_{}{}^{}\left(e_\mu c_{\mathrm{top}}(u_\mu ^{}_{d_0}^1)\right)\hfill \\ \hfill =& \phi _{\mu }^{}{}_{}{}^{}\left(e_\mu u_\mu ^{}\xi _1\mathrm{}\xi _s\underset{k=d_0l+1}{\overset{1}{}}(lH+k\mathrm{})\right).\hfill \end{array}$$
The rest is straightforward and is left to the reader. โ
## 4. Conclusion of the proofs
### 4.1. Main contribution term
<sup>7</sup><sup>7</sup>7Here instead of going through the diagrams (25) (26), it might be possible to proceed by another (equivalent) way using Giventalโs double construction formula , which reads
$$๐ข:=\underset{\beta }{}q^\beta _{[\overline{G}_{0,0}(X,\beta )]^{\mathrm{vir}}}e^{Pt}c_{\mathrm{top}}(E_\beta )=_XJ_X^E(qe^\mathrm{}t,\mathrm{})e^{pt}J_X^E(q,\mathrm{}),$$
where $`P=u_X^{}(๐ช_{_d^r}(1))`$. This should give us $`J_X^E(q,\mathrm{},t_0,t)`$ (see (36)) from formulas in Theorem 5.
Set $`\mathrm{ev}=i\mathrm{ev}_X`$ and
$$J_X^E(\beta )=J_X^E(\beta ,\text{main})+\underset{\nu }{}J_X^E(\beta ,\nu ).$$
In order to show that $`_XJ_X^E_XI_X^E`$ it is sufficient to show $`i_{}J_X^Ei_{}I_X^E`$ because
$$_Xe^{pt}\omega =_^re^{ht}i_{}\omega $$
by the projection formula.
Recall that our goal is to compute $`\mathrm{ev}_{}^{\mathrm{vir}}{\displaystyle \frac{c_{\mathrm{top}}(E_\beta )}{\mathrm{}(\mathrm{}\psi )}}`$. From (11) we have
(25)
We can easily get, by correspondence of residues (12), the โmain termโ:
$$\begin{array}{cc}\hfill i_{}J_X^E(\beta ,\text{main})=& \mathrm{ev}_{}^{\mathrm{vir}}\left(\frac{j_X^{}u_X^{}_j_{k_j}(l_jH+k_j\mathrm{})}{\mathrm{}(\mathrm{}\psi )}\right)\hfill \\ \hfill =& \frac{t^{}u_{X}^{\mathrm{vir}}{}_{}{}^{}u_X^{}_j_{k_j}(l_jH+k_j\mathrm{})}{e(N_{^r|_d^r})}\hfill \\ \hfill =& \frac{t^{}u_{X}^{\mathrm{vir}}{}_{}{}^{}1}{e(N_{^r|_d^r})}t^{}\underset{j}{}\underset{k_j}{}(l_jH+k_j\mathrm{})\hfill \\ \hfill =& \mathrm{ev}_{}^{\mathrm{vir}}(\frac{1}{\mathrm{}(\mathrm{}\psi )})\underset{j}{}\underset{k_j}{}(l_jh+k_j\mathrm{})\hfill \\ \hfill =& i_{}I_X^E.\hfill \end{array}$$
Thus $`I_X^E`$ is really the main term of $`J_X^E`$. We will see that the boundary terms are of special forms and can be taken care of by a change of variables due to the *non-negativity* condition (on the tangent bundles of complete intersection $`Y`$ in $`X`$) stated in Theorem 1.
### 4.2. Boundary terms and dimension counting
Before we start our discussion, we should remark that the term โdimensionโ here means *virtual dimension*.
Consider the commutative diagram (see (11))
(26)
$$\text{}_.$$
In the convex case (see (16)),
(27)
$$\begin{array}{cc}& i_{}J_X^E(\beta ,\nu )\hfill \\ \hfill =& \mathrm{ev}_{}^{\mathrm{vir}}\left(\frac{j_X^{}_\nu \phi _{\nu }^{}{}_{}{}^{}\frac{1}{s!}\left(e_\nu u_\nu ^{}_j_{k_j=0}^{d_0l_j}(l_jH+k_j\mathrm{})[D_\nu ]^{\mathrm{vir}}\right)}{\mathrm{}(\mathrm{}\psi )}\right)\hfill \\ \hfill =& \frac{t^{}\mathrm{PD}u_{X}^{}{}_{}{}^{}\left(_\nu \phi _{\nu }^{}{}_{}{}^{}\frac{1}{s!}\left(e_\nu u_\nu ^{}_j_{k_j=0}^{d_0l_j}(l_jH+k\mathrm{})\right)[\overline{G}_{0,0}]^{\mathrm{vir}}\right)}{_{k=1}^d(h+k\mathrm{})^{r+1}}\hfill \\ \hfill =& \frac{t^{}_\nu \frac{1}{s!}\psi _{\nu }^{}{}_{}{}^{}\left(u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu _j_{k_j=0}^{d_0l_j}(l_jH+k\mathrm{})\right)}{_{k=1}^d(h+k\mathrm{})^{r+1}}.\hfill \end{array}$$
Here we have used (13) and the left square of the above commutative diagram (27).
Similar results holds in concave case:<sup>8</sup><sup>8</sup>8We have chosen to use small font for the discussion in the convex case in this subsection.
$$\begin{array}{cc}& i_{}J_X^E(\beta ,\nu )\hfill \\ \hfill =& \mathrm{ev}_{P}^{\mathrm{vir}}{}_{}{}^{}\left(\frac{j_X^{}_\nu \phi _{\nu }^{}{}_{}{}^{}\left(e_\nu u_\nu ^{}_j_{m=1}^s\xi _m_{k_j=d_0l_j+1}^1\left(l_jH+k\mathrm{}\right)\left[D_\nu \right]^{\mathrm{vir}}\right)}{\mathrm{}\left(\mathrm{}\psi \right)}\right)\hfill \\ \hfill =& \frac{t^{}_\nu \frac{1}{s!}\psi _{\nu }^{}{}_{}{}^{}\left(u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu _j\left(\xi _1\mathrm{}\xi _s_{k_j=d_0l_j}^1\left(lH+k\mathrm{}\right)\right)\right)}{_{k=1}^d\left(h+k\mathrm{}\right)^{r+1}}.\hfill \end{array}$$
Apply the correspondence of residues (12) to the right commuting triangle of (26)
one has
(28)
$$\begin{array}{cc}& \frac{t^{}\psi _{\nu }^{}{}_{}{}^{}w}{_{k=1}^d(h+k\mathrm{})^{r+1}}\hfill \\ \hfill =& \frac{t_\nu ^{}w}{\mathrm{}^s_{k=1}^{d_0}(h+k\mathrm{})^{r+1}},\hfill \end{array}$$
Then equation (27) becomes, by (28),
(29)
$$i_{}J_X^E(\beta ,\nu )=\underset{\nu }{}\frac{_j_{k=0}^{d_0l_j}(l_jh+k\mathrm{})t_\nu ^{}u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu }{s!\mathrm{}^s_{k=1}^{d_0}(h+k\mathrm{})^{r+1}}.$$
Similarly in the concave case ($`l_j<0`$)
(30)
$$i_{}J_X^E(\beta ,\nu )=\underset{\nu }{}\frac{_j\xi _1\mathrm{}\xi _s_{k=d_0l_j+1}^1\left(l_jh+k\mathrm{}\right)t_\nu ^{}u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu }{s!\mathrm{}^s_{k=1}^{d_0}\left(h+k\mathrm{}\right)^{r+1}}.$$
*Therefore it remains to compute $`u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu `$.*
First let us work on the hypersurface case, i.e. $`E=L`$ and $`L`$ convex. Define $`L_\beta ^k:=i_G^{}(_d^k)`$ for the notational convenience. From Proposition 1
$$e_\nu =\underset{m=1}{\overset{s}{}}\underset{k=1}{\overset{d_ml}{}}(k_m)p_{\beta _m}^{}c_{\mathrm{top}}(L_{\beta _m}^{k_m+1})u_\nu ^{}c_{\mathrm{top}}(_{\mathrm{\Sigma }_m}^{k_m}/_{\mathrm{\Sigma }_m}^1)$$
where $`\mathrm{\Sigma }_m=d_0+d_1+\mathrm{}+d_{m1}`$. Again, to simplify the notations, let us start with the divisorial strata, i.e. $`\nu =(\beta _0,\beta _1)`$. In this case,
$$e_\nu =\underset{k=1}{\overset{d_1l}{}}(k)p_{\beta _1}^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})u_\nu ^{}c_{\mathrm{top}}(_{\beta _0}^k/_{\beta _0}^1).$$
Therefore
$$u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu =\underset{k=1}{\overset{d_1l}{}}(k)u_{\nu }^{\mathrm{vir}}{}_{}{}^{}p_{\beta _1}^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})c_{\mathrm{top}}(_{\beta _0}^k/_{\beta _0}^1).$$
It is easy to see that
$$c_{\mathrm{top}}(_{\beta _0}^k/_{\beta _0}^1)=\underset{m=1}{\overset{k1}{}}(l(h+d_0\mathrm{})+m\xi )$$
where $`\xi `$ is the equivariant point class of $`^1`$. Thus we only have to know $`u_{\nu }^{\mathrm{vir}}{}_{}{}^{}p_1^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})`$, which is in turn $`\mathrm{PD}u_{X,1}^{}{}_{}{}^{}p_{0}^{}{}_{}{}^{}(p_1^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})[D_\nu ]^{\mathrm{vir}})`$ because the morphism $`u_\nu :D_\nu ^1\times _d^r`$ factors through $`p_0`$
$$D_\nu \stackrel{p_0}{}G_{0,1}(X,\beta )\stackrel{u_X,1}{}^1\times _d^r.$$
For this, consider the fibre square:
(31)
so that
$$p_{0}^{}{}_{}{}^{}\left(p_1^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})[D_\nu ]^{\mathrm{vir}}\right)=\left(\mathrm{ev}_G^{}\mathrm{ev}_{M}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})\right)[\overline{G}_{0,1}(X,\beta _0)]^{\mathrm{vir}}.$$
However, by dimension counting, $`\mathrm{ev}_{M}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})`$ has the cohomological degree $`2k+c_1(L)c_1(X),\beta _1`$, which is at most one for $`k=1`$ and at most zero for $`k=2`$. It vanishes otherwise. It is also obvious that if $`c_1(L)c_1(X),\beta 2`$ for any $`\beta `$, i.e. the case Fano of index $`2`$, then all boundary contributions vanish.
From our assumption that $`i:X^r`$ induces an isomorphism $`i^{}:NS(^r)NS(X)`$, the above degree one algebraic cohomology classes lies in the image of $`i^{}`$ and we may conclude that
$$\mathrm{ev}_G^{}\mathrm{ev}_{M}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}(L_{\beta _1}^{k+1})=i_G^{}\mathrm{ev}_{\overline{G}_{0,1}(^r,d_0)}^{}(c_\nu ^0+c_\nu ^1h)$$
for $`c^i`$. However, as shown in Lemma 1 that
(32)
$$\mathrm{ev}_{\overline{G}_{0,1}(^r,d_0)}^{}h=u^{}(H+d_0\xi )ฯต,$$
where $`ฯต`$ is the exceptional divisor of $`u`$. Summing up, the boundary term for the strata $`\nu =(\beta _0,\beta _1)`$ is (see (29))
$$\begin{array}{cc}& i_{}J_X^L(\beta ,\nu )=\frac{_{k=0}^{d_0l}(lh+k\mathrm{})}{_{k=1}^d(h+k\mathrm{})}t_\nu ^{}u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu \hfill \\ \hfill =& i_{}\left(J_X(\beta _0)H_{\beta _0}^L\right)\underset{k=0}{\overset{d_0l}{}}\left(c_\nu ^0(l(h+d_0\mathrm{})+\mathrm{})+c_\nu (h_i+d_0^i\mathrm{})\right).\hfill \end{array}$$
Since $`c^0,c^i`$ are independent of $`\beta _0`$ from the above discussion, this proved the part *(a)* of the following proposition.
###### Proposition 2.
Let $`L`$ be a convex line bundle on $`X`$ induced from $`P`$, and let $`\lambda _\nu (p,\mathrm{})`$ be defined by the following formula
(33)
$$i_{}\left(J_X(\beta _0)H_{\beta _0}^L\lambda _\nu (p,\mathrm{})\right):=\frac{_{k=0}^{d_0l}(lh+k\mathrm{})t_\nu ^{}u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu }{_{k=1}^{d_0}(h+k\mathrm{})^{r+1}},$$
which is $`i_{}J_X^L(\beta ,\nu )`$.
(a) If $`\nu =(\beta _0,\beta _1)`$, then $`\lambda _\nu (p,\mathrm{})`$ is linear and satisfies
$$\lambda _{(\beta _0,\beta _1)}(p,\mathrm{})=\lambda _{(0,\beta _1)}(p+d_0\mathrm{},\mathrm{}).$$
(b) for $`\nu =(\beta _0,\beta _1,\mathrm{},\beta _s)`$, we have
(34)
$$\lambda _\nu (p,\mathrm{})=\lambda _{(\mathrm{\Sigma }_1,\beta _1)}(p,\mathrm{})\lambda _{(\mathrm{\Sigma }_2,\beta _2)}(p,\mathrm{})\mathrm{}\lambda _{(\mathrm{\Sigma }_s,\beta _s)}(p,\mathrm{}),$$
where $`\mathrm{\Sigma }_m=_{a<m}\beta _a`$.
###### Proof.
(of part *(b)*) The equation (34) requires only to compute $`t_\nu ^{}u_{\nu }^{}{}_{}{}^{}e_\nu `$, with $`e_\nu `$ described by (29)
$$u_{\nu }^{}{}_{}{}^{}e_\nu =\underset{m}{}\underset{k_m}{}(k_m)u_{\nu }^{}{}_{}{}^{}p_{\beta _m}^{}c_{\mathrm{top}}(L_{\beta _m}^{k_m+1})c_{\mathrm{top}}(_{\mathrm{\Sigma }_m}^{k_m}/_{\mathrm{\Sigma }_m}^1).$$
Using the same argument to the following diagram
(35)
one obtains that
$$c_{\mathrm{top}}(_{\mathrm{\Sigma }_m}^{k_m}/_{\mathrm{\Sigma }_m}^1)=\underset{a=1}{\overset{k_m1}{}}\left(lh+l\mathrm{\Sigma }_m\mathrm{}+\underset{b=1}{\overset{m1}{}}ld_b\xi _b+a\xi _m\right).$$
Dimension counting and (32) gives us the proof of part *(b)*. โ
In the concave case a similar modification goes through.
###### Proposition 3.
Let $`L`$ be a concave line bundle and let $`\lambda _\nu `$ be defined as in (33). Then part (a) and (b) hold. Furthermore, $`\lambda _\nu `$ depends only on $`\mathrm{}`$.
###### Proof.
From (30) for $`J_P^L(\beta ,\nu )`$, we only have to compute $`u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu `$. A straightforward modification, by fibre square (31) and dimension counting, will lead to the conclusion that only degree zero (constant) terms survives in $`\mathrm{ev}_{M}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}\left(L_{\beta _1}^{k+1}\right)`$. The difference is that the cohomological degree of $`\mathrm{ev}_{M}^{\mathrm{vir}}{}_{}{}^{}c_{\mathrm{top}}\left(L_{\beta _1}^{k+1}\right)`$ is only $`k+c_1\left(L\right)c_1\left(X\right),\beta _1`$, because the rank of $`L^{k+1}`$ in this case is one less than that of $`L^{k+1}`$ in convex case.
Therefore $`\lambda _\nu =constantt_\nu ^{}_{m=1}^s\xi _m=constant\mathrm{}^s`$. This allows us to set $`\lambda _{(\beta _0,\beta _1)}=constant\mathrm{}`$. โ
###### Corollary 5.
The same is true when the rank of $`E`$ is greater than 1.
The proof uses the inductive property of $`e_\nu `$ and is completely analogous to the above argument.
Summarizing the above discussion, we have
###### Theorem 6.
(36)
$$i_{}J_X^E=i_{}\underset{\beta }{}q^\beta \underset{\nu }{}\frac{J_X(\beta _0)H_{\beta _0}^E_{m=1}^s\lambda _{\beta _m}(p+\mathrm{\Sigma }_m\mathrm{},\mathrm{})}{s!\mathrm{}^s}$$
such that $`\lambda _{\beta _m}(p,\mathrm{})`$ are linear form in $`p,\mathrm{}`$ and is independent of the total degree $`\beta `$ in the background.
###### Proof.
Set $`\lambda _{\beta _m}(p,\mathrm{})=\lambda _{(0,\beta _m)}(p,\mathrm{})`$ and apply Propositions 2 and 3. โ
### 4.3. Mirror transformation
*In this subsection we will omit the push-forward symbol $`i_{}`$. All equalities are assumed to hold after pushing-forward to $`^r`$.*
Now we are ready to find a change of variables (mirror transformation) between $`J_X^E`$ and $`I_X^E`$. Note that we will use $`\beta `$ both as an element in $`H_2(X)`$ and a number. In the case $`Pic(X)=1`$, this should not cause confusion.
$$\begin{array}{cc}& t_0t_0+f_0(q)\mathrm{},\hfill \\ & tt+f_1(q),\hfill \end{array}$$
where $`f_0(q),f_1(q)`$ are formal power series in $`q`$. After the change of variables, $`e^{(t_0+pt)/\mathrm{}}I_X^E(q,\mathrm{})`$ becomes
(37)
$$e^{f_0(q)+\frac{1}{\mathrm{}}pf_1(q)}e^{\frac{1}{\mathrm{}}(t_0+pt)}I_X^E(qe^{f_1(q)},\mathrm{}),$$
*which we would like to equate to $`J_X^E(q,\mathrm{})`$*.
Let us set $`f_1(q)=_\beta a_\beta q^\beta `$ and $`f_0(q)=_\beta b_\beta q^\beta `$. Then (37) can be expanded as:
(38)
$$\begin{array}{cc}& e^{f_0(q)+\frac{1}{\mathrm{}}hf_1(q)}e^{\frac{1}{\mathrm{}}(t_0+pt)}I_X^E(qe^{f_1(q)},\mathrm{})\hfill \\ \hfill =& \underset{\beta _0}{}q^{\beta _0}e^{f_0(q)+(\frac{p}{\mathrm{}}+\beta _0)f_1(q)}I_X^E(\beta _0)\hfill \\ \hfill =& \underset{\beta }{}q^\beta \underset{_{m=0}^s\beta _m=\beta }{}\frac{1}{s!}I_X^E(\beta _0)\underset{m=1}{\overset{s}{}}\left(b_{\beta _m}+a_{\beta _m}(\beta _0+\frac{p}{\mathrm{}})\right).\hfill \end{array}$$
In order to prove Theorem 1 we need to find $`a_\beta ,b_\beta `$ so that (38) is equal to $`J_X^E(q,\mathrm{})`$. The following simple lemma in is useful:
###### Lemma 2.
Let
(39)
$$Q(q)=\underset{\beta }{}q^\beta \underset{_{m=1}^s\beta _m=\beta ,d_m0}{}\frac{1}{s!}\underset{m=0}{\overset{s}{}}\left(y_{\beta _m}+x_{\beta _m}B_{m1}\right),$$
where $`B_{m1}:=_{k=1}^{m1}\beta _k`$, $`(B_0=0)`$. Then $`\mathrm{log}(Q(q))`$ is a linear function of $`y_\beta `$.
###### Corollary 6.
Use the notation from the above lemma. Define $`z_\beta `$ by
(40)
$$Q(q)=\mathrm{exp}(\underset{\beta 0}{}z_\beta q^\beta ).$$
Then
(41)
$$z_\beta =\underset{_{m=1}^s\beta _m=\beta }{}\frac{1}{s!}y_{\beta _1}\underset{m=1}{\overset{s}{}}x_{\beta _m}B_{m1}$$
###### Proof.
Expand equations (40) and (39). The $`q^\beta `$ term is
(42)
$$\underset{_{m=1}^s\beta _m=\beta }{}\frac{1}{s!}\underset{m=1}{\overset{s}{}}z_{\beta _m}=\underset{_{m=1}^s\beta _m=\beta }{}\frac{1}{s!}\underset{m=1}{\overset{s}{}}(y_{\beta _m}+x_{\beta _m}B_{m1}).$$
By Lemma 2 the right hand side of (42) is linear with respect to $`y_\beta `$. The linear in $`y_\beta `$ term on the RHS is exactly (41). โ
Back to the proof of Theorem 1. We wish to equate (36) to (38). Set $`\lambda _\beta =c_\beta ^0\mathrm{}+c_\beta ^1p`$, then (36) can be written as
$$\begin{array}{cc}\hfill J_X^E(q)=& \underset{\beta _0}{}\underset{_{m=1}^s\beta _m^{}=\beta ^{}}{}q^{\beta ^{}+\beta _0}J_X(\beta _0)\left(\underset{j}{}H_{\beta _0}^{L_j}\right)\hfill \\ & \frac{1}{s!}\underset{m=1}{\overset{s}{}}\left(\left(c_{\beta _m^{}}^0+c_{\beta _m^{}}^1(\frac{p}{\mathrm{}}+\beta _0)\right)+c_{\beta _m^{}}^1B_{m1}\right)\hfill \\ \hfill =& \underset{\beta _0}{}q^{\beta _0}J_X(\beta _0)\left(\underset{j}{}H_{\beta _0}^{L_j}\right)\hfill \\ & \underset{_{m=1}^s\beta _m^{}=\beta ^{}}{}q^\beta ^{}\frac{1}{s!}\underset{m=1}{\overset{s}{}}\left(\left(c_{\beta _m^{}}^0+c_{\beta _m^{}}^1(\frac{p}{\mathrm{}}+\beta _0)\right)+c_{\beta _m^{}}^1B_{m1}\right).\hfill \end{array}$$
Apply Corollary 6, we find a new variables $`z_\beta ^{}`$ such that $`z_\beta ^{}`$ are linear in $`c_{\beta _m^{}}^0+c_{\beta _m^{}}^1(\frac{p}{\mathrm{}}+\beta _0)`$ and polynomial in $`c_{\beta _j^{}}^1`$. Therefore there are constants $`b_\beta ^{},a_\beta ^{}`$ such that $`z_\beta ^{}=b_\beta ^{}+a_\beta ^{}(\frac{p}{\mathrm{}}+\beta _0)`$, and $`b_\beta ^{},a_\beta ^{}`$ are independent of $`\beta _0`$. This implies that
$$J_X^E(q)=\underset{_{m=0}^s\beta _m=\beta }{}q^\beta I_X^E(\beta _0)\frac{1}{s!}\underset{m=1}{\overset{s}{}}\left(b_{\beta _m}+a_{\beta _m}(\frac{p}{\mathrm{}}+\beta _0)\right),$$
which is exactly (38). One can easily see from the above proof that the case $`E`$ being a direct sum of convex and concave bundles requires little modification. Our proof of the Theorem 1 is therefore complete.
### 4.4. Proof of Theorem 2
The proof of case 2 is very easy. As we have seen in ยง 4.2 that the use of fibre square (31) (35) and dimension counting guarantees that $`u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu =0`$. Therefore $`J_X^E(\beta ,\nu )=0`$.
The proof of case 1 goes a slightly different way.<sup>9</sup><sup>9</sup>9It is also possible to prove the convex case in this way. . As we have seen in the proof of Proposition 3 that $`u_{\nu }^{\mathrm{vir}}{}_{}{}^{}e_\nu `$ contains only constant terms. Therefore the boundary contribution should come from
$$\overline{G}_{0,s}(X,\beta _0)\stackrel{u_s}{}(^1)^s\times P_{d_0}\stackrel{\psi _\nu }{}_d^r,$$
and should be of the form
$$\psi _{\nu }^{}{}_{}{}^{}\left((constant)(u_{s}^{}{}_{}{}^{}1)\underset{j=1}{\overset{\mathrm{rk}(E)}{}}\xi _1\mathrm{}\xi _s\underset{k_j=c_1(L_j),\beta _0+1}{\overset{1}{}}(l_jH+k_j\mathrm{})\right),$$
which has cohomological dimension greater than $`J_X^E(\beta ,\text{main})`$. Therefore $`constant=0`$.
The proof of direct sum case is the combination of the above arguments.
### 4.5. Proof of Corollary 2
To simplify our notations, set
$$J^{}:=i_{}c_{\mathrm{top}}(E)e^{\frac{1}{\mathrm{}}(t_0+pt)}J_X^E(q,\mathrm{}),$$
and
$$J_{}^{}:=i_{}\frac{1}{c_{\mathrm{top}}(E)}e^{\frac{1}{\mathrm{}}(t_0+pt)}J_X^E^{}(q,\mathrm{}).$$
Notice that $`J_{}^{}`$ is actually well defined as $`J_X^E^{}(q,\mathrm{})`$ always has a factor $`c_{\mathrm{top}}(E^{})`$. More precisely, consider the exact sequence
$$0R^0\pi _{}(\mathrm{ev}^{}(E^{})๐ช(x_1))R^0\pi _{}\mathrm{ev}^{}(E^{})\mathrm{ev}^{}E^{}0.$$
Then in fact
$$\begin{array}{cc}\hfill J_{}^{}=& i_{}e^{\frac{1}{\mathrm{}}(t_0+pt)}\underset{\beta }{}q^\beta \mathrm{ev}_{}^{\mathrm{vir}}\left(\frac{c_{\mathrm{top}}\left(R^0\pi _{}\left(\mathrm{ev}^{}(E^{})๐ช(x_1)\right)\right)}{\mathrm{}(\mathrm{}\psi )}\right)\hfill \\ & =i_{}e^{\frac{1}{\mathrm{}}(t_0+pt)}J_X^E^{}\frac{1}{c_{\mathrm{top}}(E^{})}.\hfill \end{array}$$
Similar interpretation holds for $`J^{}`$: it is the $`J`$-function of the bundle $`R^1\pi _{}(\mathrm{ev}^{}(E)๐ช(x_1))`$, multiplied by $`e^{\frac{1}{\mathrm{}}(t_0+pt)}`$. Notice that the ranks of the two bundles $`R^0\pi _{}(\mathrm{ev}^{}(E^{})๐ช(x_1))`$ and $`R^1\pi _{}(\mathrm{ev}^{}(E)๐ช(x_1))`$ are the same.
It follows from Theorems 1 and 2
$$J^{}=I^{}:=e^{\frac{1}{\mathrm{}}(t_0+pt)}\underset{\beta }{}q^\beta J_X(\beta )\underset{j}{}\underset{k_j=1}{\overset{c_1(L_j),\beta }{}}(c_1(L)+k\mathrm{})$$
and
$$J_{}^{}I_{}^{}:=(1)^{\mathrm{rk}(E)}e^{\frac{1}{\mathrm{}}(t_0+pt)}\underset{\beta }{}q^\beta J_X(\beta )\underset{j}{}\underset{k_j=c_1(L_j),\beta +1}{\overset{0}{}}(c_1(L)+k\mathrm{}).$$
Notice the change of limits in the products of $`k`$ due to $`c_{\mathrm{top}}(E)`$ and $`c_{\mathrm{top}}(E^{})`$ and the possible sign coming from the ratio of $`c_{\mathrm{top}}(E)`$ and $`c_{\mathrm{top}}(E^{})`$.
The difference between $`I^{}`$ and $`I_{}^{}`$ can be easily figured out by using the quantum differential equation .
###### Claim.
There is a power series $`\varphi (q)`$ such that
(43)
$$\varphi (q)I^{}=I_{}^{}.$$
We now sketch the proof of (43). Fix a basis $`\{e_a\}`$ of the vector space $`H^{}(X)`$ with $`e_1=1`$. Let
$$S_{ab}:=\underset{\beta }{}q^\beta (e_a,e^{\frac{1}{\mathrm{}}(t_0+pt)}\frac{e_b}{\mathrm{}\psi })_{0,2,\beta }$$
be a matrix of genus zero two-point GromovโWitten invariants via the โfundamental classโ $`c_{\mathrm{top}}(E_\beta )c_{\mathrm{top}}(E)[\overline{M}_{0,2}(X,\beta )]^{\mathrm{vir}}`$. Note that we have identified $`t_i=\mathrm{log}q_i`$. Giventalโs theory of quantum differential equation says
(44)
$$\mathrm{}\frac{}{t_i}S=p_iS,$$
where $`p_i`$ is the quantum multiplication matrix. By the (virtual) dimension counting $`J_X`$ is a polynomial in $`\mathrm{}^1`$ of the form $`1+O(\mathrm{}^{c_1(E),\beta })`$. Therefore
$$I^{}=J^{}=(1+O(\mathrm{}^{\mathrm{rk}(E)}))e^{\frac{1}{\mathrm{}}(t_0+pt)}.$$
Notice that $`S_{1a}=J^{},e_a`$. This implies that the first row $`S_{1a}`$ of $`S`$ has the same order in $`\mathrm{}^1`$.
Set $`v_j:=c_1(L_j)`$ and $`_j`$ be the directional derivative on the direction $`v_j`$. First of all $`(_{j=1}^{\mathrm{rk}(E)}_j)(I^{}/c_{\mathrm{top}}(E))=I_{}^{}`$ by simple derivations. Now apply induction on the order $`m`$ of the differential operator $`๐_m:=_{j=1}^m\mathrm{}_j`$. I claim that
(45)
$$๐_m(S)=\left(\underset{j=1}{\overset{m}{}}v_j\right)S$$
for $`m<\mathrm{rk}(E)`$.
The case $`m=1`$ is true by (44). Suppose that (45) is true for some $`m<\mathrm{rk}(E)1`$. This implies that $`๐_mS=(_j^mv)+O(\mathrm{}^1)`$. Therefore $`((_j^mv))_{1a}`$ are independent of $`t`$ from LHS. Now differentiate one more time:
(46)
$$๐_{m+1}S=(\mathrm{}_{m+1}\left(\underset{j=1}{\overset{m}{}}v_j\right))S+(\left(\underset{j=1}{\overset{m}{}}v_j\right)v_{m+1})S.$$
and the first row of RHS will be $`(_{j=1}^mv_j)v_{m+1},e_a+O(\mathrm{}^1)`$. Because the first row of the LHS modulo $`\mathrm{}^1`$ is independent of $`t`$, we have again $`(_{j=1}^mv_j)v_{m+1}=_{j=1}^{m+1}v_j`$. Thus (45) holds. If $`m=\mathrm{rk}(E)1`$, then the first row of LHS of (46) would be certain power series $`\varphi (q)`$ of $`q`$ (mod $`\mathrm{}^1`$). Since the first row of first term on RHS still vanishes (mod $`\mathrm{}^1`$), we have the first row of (46) equal to
$$\begin{array}{cc}\hfill \varphi (q)I^{}=& \varphi (q)\left(\underset{j=1}{\overset{\mathrm{rk}(E)}{}}v_j\right)\left(\frac{I^{}}{c_{\mathrm{top}}(E)}\right)\hfill \\ \hfill =& \left(\underset{j=1}{\overset{\mathrm{rk}(E)}{}}(v_j)\right)I^{}=๐_{\mathrm{rk}(E)}I^{}\hfill \\ \hfill =& \pm I_{}^{}.\hfill \end{array}$$ |
warning/0003/math-ph0003025.html | ar5iv | text | # 1 INTRODUCTION
## 1 INTRODUCTION
The oscillator algebra of creation, annihilation, and number operators plays a central role in the investigation of many physical systems, and provides a useful tool in the theory of Lie algebra representations. Similarly, some of its deformations (or extensions) have found applications to various physical problems, such as the description of systems with non-standard statistics (Greenberg, 1990, 1991; Fivel, 1990; Meljanac et al., 1994; Meljanac and Milekoviฤ, 1996; Quesne, 1994a), the construction of integrable lattice models (Bogoliubov et al., 1994), the investigation of nonlinearities in quantum optics (McDermott and Solomon, 1994; Solomon, 1998; Manโko et al., 1997), the bosonization of supersymmetric quantum mechanics (SSQM) (Bonatsos and Daskaloyannis, 1993a; Brzeziลski et al., 1993; Plyushchay, 1996a, b; Beckers et al., 1997), as well as the algebraic treatment of quantum exactly solvable models (Daskaloyannis, 1992; Bonatsos and Daskaloyannis, 1993b; Bonatsos et al., 1993, 1994; Quesne, 1994b), $`n`$-particle integrable systems (Vasiliev, 1991; Polychronakos, 1992; Brink et al., 1992; Brink and Vasiliev, 1993; Quesne, 1995), pairing correlations in nuclei (Bonatsos, 1992; Bonatsos and Daskaloyannis, 1992a), and vibrational spectra of molecules (Chang et al., 1991; Chang and Yan, 1991a, b, c; Bonatsos and Daskaloyannis, 1992b, 1993c). In addition, they have been used to construct representations of quantum universal enveloping algebras of Lie algebras, also referred to as quantum algebras (Biedenharn, 1989; Macfarlane, 1989; Sun and Fu, 1989; Hayashi, 1990; Fairlie and Zachos, 1991; Fairlie and Nuyts, 1994).
Deformations of the oscillator algebra arose from successive generalizations of the Arik-Coon (Arik and Coon, 1976; Kuryshkin, 1980), and Biedenharn-Macfarlane (Biedenharn, 1989; Macfarlane, 1989; Sun and Fu, 1989) $`q`$-oscillators. Various attempts have been made to introduce some order in the rich and varied choice of deformed commutation relations by defining โgeneralized deformed oscillator algebrasโ (GDOAs). Among them, one may quote the treatments due to Jannussis et al. (1991), Jannussis (1993), Daskaloyannis (1991), Bonatsos and Daskaloyannis (1993a), Irac-Astaud and Rideau (1992, 1993, 1994), McDermott and Solomon (1994), Meljanac et al. (1994), Meljanac and Milekoviฤ (1996), Katriel and Quesne (1996), Quesne and Vansteenkiste (1995, 1996, 1997). In the remainder of the present paper, we shall refer to GDOAs as defined in the last references.
$`G`$-extended oscillator (or alternatively Heisenberg<sup>1</sup>) algebras, where $`G`$ is some finite group, appeared in connection with $`n`$-particle integrable models. It was shown (Vasiliev, 1991; Polychronakos, 1992; Brink et al., 1992; Brink and Vasiliev, 1993; Quesne, 1995) that they provide an algebraic formulation of the Calogero model (Calogero, 1969a, b, 1971), or some generalization thereof (Wolfes, 1974; Calogero and Marchioro, 1974). In the former case, $`G`$ is the symmetric group $`S_n`$ (Polychronakos, 1992; Brink et al., 1992; Brink and Vasiliev, 1993). For two particles, the abelian group $`S_2`$ can be realized in terms of Klein operator $`K=(1)^N`$, where $`N`$ denotes the number operator. The $`S_2`$-extended oscillator algebra then becomes a GDOA, also known as the Calogero-Vasiliev (Vasiliev, 1991), or modified (Brzeziลski et al., 1993) oscillator algebra. Some deformations of the latter have been extensively studied (Brzeziลski et al., 1993; Macfarlane, 1994; Kosiลski et al., 1997; Tsohantjis et al., 1997; Paolucci and Tsohantjis, 1997).
The purpose of the present paper is to study a new class of $`G`$-extended oscillator algebras (Quesne and Vansteenkiste, 1998), generalizing the one describing the two-particle Calogero model. Here $`G`$ is the cyclic group of order $`\lambda `$, $`C_\lambda =\{I,T,T^2,\mathrm{},T^{\lambda 1}\}`$, which for $`\lambda =2`$ is isomorphic to $`S_2`$. Such $`C_\lambda `$-extended oscillator algebras $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ have a rich structure, since they depend upon $`\lambda 1`$ independent real parameters $`\alpha _0`$, $`\alpha _1`$, $`\mathrm{}`$$`\alpha _{\lambda 2}`$ (reducing to a single one in the $`\lambda =2`$ case, corresponding to the $`S_2`$-extended oscillator algebra). Realizing $`T`$ in terms of the number operator $`N`$ converts $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ into a GDOA $`๐^{(\lambda )}(G(N))`$.
The bosonic oscillator Hamiltonian $`H_0`$, associated with $`๐^{(\lambda )}(G(N))`$, is equivalent to the two-particle Calogero Hamiltonian for $`\lambda =2`$, but exhibits entirely new features for $`\lambda 3`$ (Quesne and Vansteenkiste, 1998). In such a case, all the levels corresponding to a number of quanta equal to $`\mu \text{mod}\lambda `$ are equally spaced, but the ordering and spacing of levels associated with different $`\mu `$ values depend on the algebra parameters $`\alpha _0`$, $`\alpha _1`$, $`\mathrm{}`$$`\alpha _{\lambda 2}`$. By appropriately choosing the latter, one may therefore obtain nondegenerate spectra, as well as spectra exhibiting some $`(\nu +1)`$-fold degeneracies, where $`\nu `$ may take any value in the set $`\{1,2,\mathrm{},\lambda 1\}`$.
The rich variety of spectra that may be obtained with $`H_0`$, as well as the connection with the Calogero model for $`\lambda =2`$, makes it most likely that some interesting applications will arise in one or another context. To help towards finding them, the construction of realizations of the $`๐^{(\lambda )}(G(N))`$ generators in terms of differential operators is under current investigation and will be reported elsewhere.
We may however already note that spectra that are a strictly equidistant continuation of a triplet of โgroundโ states (which can be obtained here for $`\lambda =3`$) arose in two studies of a class of potentials (with applications in string theory) using either an advanced factorization method (Veselov and Shabat, 1993), or a nonlinear generalization of the Fock method (Eleonsky et al., 1994, 1995; Eleonsky and Korolev, 1995). Such spectra can also be obtained in SSQM by using cyclic shape invariant potentials of period three (Sukhatme et al., 1997). In this context, we recently showed that three appropriately chosen $`๐^{(3)}(G(N))`$ algebras provide a matrix realization of SSQM (Quesne and Vansteenkiste, 1999).
Another field wherein $`C_\lambda `$-extended oscillator algebras and their deformations may be of interest is the study of coherent (or squeezed) states in nonlinear quantum optics, wherein nonlinear oscillators are known to play an important role (McDermott and Solomon, 1994; Solomon, 1998; Manโko et al., 1997).
In the present paper, apart from studying some mathematical properties of $`C_\lambda `$-extended oscillator algebras, we deal with some important conceptual applications of these algebras. We indeed plan to show that they provide a bosonization (i.e., a realization in terms of only boson-like operators without fermion-like ones) of several variants of SSQM, namely parasupersymmetric quantum mechanics (PSSQM) of arbitrary order $`p`$ (Rubakov and Spiridonov, 1988; Khare, 1992, 1993), pseudoSSQM (Beckers et al., 1995a, b; Beckers and Debergh, 1995a, b), and orthosupersymmetric quantum mechanics (OSSQM) of order two (Khare et al., 1993a). These results generalize that previously obtained for standard SSQM in terms of the Calogero-Vasiliev algebra (Brzeziลski et al., 1993; Plyushchay, 1996a, b).
In Section 2, we review the definition of $`C_\lambda `$-extended oscillator algebras, give their Casimir operators, and present some of their realizations. In Section 3, we classify their unitary irreducible representations (unirreps). In Sections 4, 5, and 6, we consider their applications to PSSQM of arbitrary order $`p`$, pseudoSSQM, and OSSQM of order two, respectively. In Section 7, we construct some of their deformations. Finally, Section 8 contains the conclusion.
## 2 $`๐ช_๐`$-EXTENDED OSCILLATOR ALGEBRAS
A $`C_\lambda `$-extended oscillator algebra $`๐^{(\lambda )}`$, where $`\lambda `$ may take any value in the set $`\{2,3,4,\mathrm{}\}`$, is defined (Quesne and Vansteenkiste, 1998) as an algebra generated by the operators $`I`$, $`a^{}`$, $`a=\left(a^{}\right)^{}`$, $`N=N^{}`$, and $`T=\left(T^{}\right)^1`$, satisfying the relations
$$[N,a^{}]=a^{},[N,T]=0,T^\lambda =I,$$
(2.1)
$$[a,a^{}]=I+\underset{\mu =1}{\overset{\lambda 1}{}}\kappa _\mu T^\mu ,a^{}T=e^{\mathrm{i2}\pi /\lambda }Ta^{},$$
(2.2)
together with their Hermitian conjugates. Here $`\kappa _\mu `$, $`\mu =1`$, 2, $`\mathrm{}`$$`\lambda 1`$, are some complex parameters restricted by the conditions $`\kappa _\mu ^{}=\kappa _{\lambda \mu }`$ (so that there remain altogether $`\lambda 1`$ independent real parameters), and $`T`$ is the generator of the cyclic group of order $`\lambda `$, $`C_\lambda =\{I,T,T^2,\mathrm{},T^{\lambda 1}\}`$ (or, more precisely, the generator of a unitary representation thereof).
It is well known (Cornwell, 1984) that $`C_\lambda `$ has $`\lambda `$ inequivalent, one-dimensional matrix unirreps $`\mathrm{\Gamma }^\mu `$, $`\mu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, which are such that $`\mathrm{\Gamma }^\mu \left(T^\nu \right)=\mathrm{exp}(\mathrm{i2}\pi \mu \nu /\lambda )`$ for any $`\nu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$. The projection operator on the carrier space of $`\mathrm{\Gamma }^\mu `$ may be written as
$$P_\mu =\frac{1}{\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}\left(\mathrm{\Gamma }^\mu \left(T^\nu \right)\right)^{}T^\nu =\frac{1}{\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu \nu /\lambda }T^\nu ,$$
(2.3)
and conversely $`T^\nu `$, $`\nu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, may be expressed in terms of the $`P_\mu `$โs as
$$T^\nu =\underset{\mu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu \nu /\lambda }P_\mu .$$
(2.4)
The algebra defining relations (2.1) and (2.2) may therefore be rewritten in terms of $`I`$, $`a^{}`$, $`a`$, $`N`$, and $`P_\mu ^{}=P_\mu ^{}`$, $`\mu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, as
$$[N,a^{}]=a^{},[N,P_\mu ]=0,a^{}P_\mu =P_{\mu +1}a^{},$$
(2.5)
$$\underset{\mu =0}{\overset{\lambda 1}{}}P_\mu =I,P_\mu P_\nu =\delta _{\mu ,\nu }P_\mu ,$$
(2.6)
$$[a,a^{}]=I+\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,$$
(2.7)
where we use the convention $`P_\mu ^{}=P_\mu `$ if $`\mu ^{}\mu =0\mathrm{mod}\lambda `$ (and similarly for other operators or parameters labelled by $`\mu `$, $`\mu ^{}`$). Equations (2.5)โ(2.7) depend upon $`\lambda `$ real parameters $`\alpha _\mu `$, $`\mu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, defined in terms of the $`\kappa _\mu `$โs by
$$\alpha _\mu =\underset{\nu =1}{\overset{\lambda 1}{}}\mathrm{exp}(\mathrm{i2}\pi \mu \nu /\lambda )\kappa _\nu ,\mu =0,1,\mathrm{},\lambda 1,$$
(2.8)
and restricted by the condition
$$\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu =0.$$
(2.9)
Hence, we may eliminate one of them, e.g., $`\alpha _{\lambda 1}`$, and denote the algebra by $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$. It will, however, often prove convenient to work instead with the $`\lambda `$ dependent parameters $`\alpha _0`$, $`\alpha _1`$, $`\mathrm{}`$$`\alpha _{\lambda 1}`$.
From Eqs. (2.1) and (2.2), or (2.5)โ(2.7), it is easy to check that $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ admits the following Casimir operators:
$$๐_1=e^{\mathrm{i2}\pi N},$$
(2.10)
$$๐_2=e^{\mathrm{i2}\pi N/\lambda }T=\underset{\mu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi (N\mu )/\lambda }P_\mu ,$$
(2.11)
$$๐_3=N+\underset{\mu =0}{\overset{\lambda 1}{}}\beta _\mu P_\mu a^{}a,$$
(2.12)
where
$$\beta _\mu =\underset{\nu =0}{\overset{\mu 1}{}}\alpha _\nu ,\mu =1,2,\mathrm{},\lambda 1,$$
(2.13)
and $`\beta _0=\beta _\lambda =0`$. The first two operators are not functionally independent since
$$๐_1๐_2^\lambda =I.$$
(2.14)
From Eq. (2.11), it follows that the cyclic group generator $`T`$ can be rewritten in terms of $`N`$ and $`๐_2`$ as
$$T=e^{\mathrm{i2}\pi N/\lambda }๐_2.$$
(2.15)
The simplest realization of the cyclic group $`C_\lambda `$ uses functions of $`N`$. By taking $`๐_2=I`$ in Eq. (2.15), and using Eq. (2.3), we obtain
$$T=e^{\mathrm{i2}\pi N/\lambda },P_\mu =\frac{1}{\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \nu (N\mu )/\lambda },\mu =0,1,\mathrm{},\lambda 1.$$
(2.16)
With such a choice, $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ becomes a GDOA $`๐^{(\lambda )}(G(N))`$, i.e., an algebra generated by $`I`$, $`a^{}`$ , $`a=\left(a^{}\right)^{}`$, and $`N=N^{}`$, subject to the relations
$$[N,a^{}]=a^{},[a,a^{}]=G(N),$$
(2.17)
where $`G(N)`$ is some Hermitian, analytic function of $`N`$ (Quesne and Vansteenkiste, 1995). In the present case,
$$G(N)=I+\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,$$
(2.18)
where $`P_\mu `$ is given by Eq. (2.16).
According to the GDOA general theory (see Quesne and Vansteenkiste (1995, 1996, 1997) and references quoted therein), one may define a structure function $`F(N)`$, which is the solution of the difference equation $`F(N+1)F(N)=G(N)`$ such that $`F(0)=0`$. For $`G(N)`$ given in Eq. (2.18), one finds
$$F(N)=N+\underset{\mu =0}{\overset{\lambda 1}{}}\beta _\mu P_\mu ,$$
(2.19)
where $`\beta _\mu `$ is defined in Eq. (2.13). From Eq. (2.19), it follows that the two Casimir operators $`๐_1`$, $`๐_3`$ of Eqs. (2.10), (2.12) reduce to the well-known Casimir operators, $`U=\mathrm{exp}(\mathrm{i2}\pi N)`$ and $`๐=F(N)a^{}a`$, respectively (Quesne and Vansteenkiste, 1996, 1997).
It is worth noting that there exist other realizations of $`C_\lambda `$, which may be interesting in some physical applications. We shall mention here two of them.
The first one uses functions of spin $`s`$ operators, where $`s=(\lambda 1)/2`$. Denoting as usual the spin operators (generating an su(2) Lie algebra) by $`S_i`$, $`i=1`$, 2, 3, it is obvious that the operators
$$P_\mu =\underset{\genfrac{}{}{0pt}{}{\sigma =\left(\lambda 1\right)/2}{\sigma \left(\lambda 2\mu 1\right)/2}}{\overset{(\lambda 1)/2}{}}\frac{S_3\sigma }{\frac{1}{2}(\lambda 2\mu 1)\sigma },\mu =0,1,\mathrm{},\lambda 1,$$
(2.20)
acting in spin space, project on the spin components $`\sigma =(\lambda 1)/2`$, $`(\lambda 3)/2`$, $`\mathrm{}`$, $`(\lambda 2\mu 1)/2`$, $`\mathrm{}`$, $`(\lambda 1)/2`$, respectively. The corresponding realization of the $`C_\lambda `$ generator $`T`$ is obtained from Eq. (2.4) in the form
$$T=\underset{\mu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu /\lambda }\left(\underset{\genfrac{}{}{0pt}{}{\nu =0}{\nu \mu }}{\overset{\lambda 1}{}}\frac{2S_3\lambda +2\nu +1}{2(\nu \mu )}\right).$$
(2.21)
By using the $`(2s+1)\times (2s+1)`$ matrix representation of $`S_3`$, $`S_3=\mathrm{diag}(s,s1,\mathrm{},s)`$, we get another realization of $`C_\lambda `$ in terms of $`\lambda \times \lambda `$ matrices,
$$T=\underset{\mu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu /\lambda }e_{\mu +1,\mu +1},P_\mu =e_{\mu +1,\mu +1},$$
(2.22)
where $`e_{ij}`$ denotes the $`\lambda \times \lambda `$ matrix with 1 in row $`i`$ and column $`j`$, and zeros everywhere else.
Note that when considering such realizations of $`C_\lambda `$, the remaining $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ generators would either act in both configuration and spin spaces, or be $`\lambda \times \lambda `$ operator-valued matrices.
For $`\lambda =2`$, the last relation in Eq. (2.1) and those in Eq. (2.2) become
$$T^2=I,\{a^{},T\}=0,[a,a^{}]=I+\kappa _1T=I+\alpha _0(P_0P_1),$$
(2.23)
where $`P_0=(I+T)/2`$, $`P_1=(IT)/2`$, and $`\kappa _1`$, $`\alpha _0\text{}`$. In the corresponding GDOA, the operator $`T`$ is given by $`T=\mathrm{exp}(\mathrm{i}\pi N)`$, which amounts to Klein operator $`K=(1)^N`$, since as shown in the next section, the eigenvalues of $`N`$ are integer in the $`๐^{(2)}(G(N))`$ unirreps. In the matrix realization (2.22), T is represented by the Pauli spin matrix $`\sigma _3`$, while $`a^{}`$, $`a`$ can be expressed in terms of $`\sigma _1`$, $`\sigma _2`$, and some differential operators (Bagchi, 1994).
For $`\lambda =3`$, the counterpart of Eq. (2.23) reads
$$T^3=I,a^{}T=e^{\mathrm{i2}\pi /3}Ta^{},$$
(2.24)
$$[a,a^{}]=I+\kappa _1T+\kappa _1^{}T^2=I+\alpha _0P_0+\alpha _1P_1(\alpha _0+\alpha _1)P_2,$$
(2.25)
where $`P_0=(I+T+T^2)/3`$, $`P_1=\left(I+e^{\mathrm{i2}\pi /3}T+e^{\mathrm{i4}\pi /3}T^2\right)/3`$, $`P_2=\left(I+e^{\mathrm{i4}\pi /3}T+e^{\mathrm{i2}\pi /3}T^2\right)/3`$, $`\kappa _1\text{}`$, and $`\alpha _0`$, $`\alpha _1\text{}`$. In the GDOA realization, the operator $`T`$ is given by $`T=\mathrm{exp}(\mathrm{i2}\pi N/3)`$, so that $`G(N)=I+2(\mathrm{}e\kappa _1)\mathrm{cos}(2\pi N/3)2(\mathrm{}m\kappa _1)\mathrm{sin}(2\pi N/3)`$. In the matrix realization (2.22), $`T`$ is represented by the matrix $`\mathrm{diag}(1,e^{\mathrm{i2}\pi /3},e^{\mathrm{i4}\pi /3})`$. Explicit expressions of $`a^{}`$, $`a`$ are still unknown.
In the remainder of this paper, we shall concentrate on the abstract definition of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, or its GDOA realization $`๐^{(\lambda )}(G(N))`$.
## 3 UNIRREPS OF $`๐ช_๐`$-EXTENDED OSCILLATOR ALGEBRAS
The purpose of the present section is to provide a classification of the $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ unirreps. To carry out this program, it proves convenient to first consider the corresponding GDOA $`๐^{(\lambda )}(G(N))`$, defined in Eqs. (2.16)โ(2.18).
### 3.1 Unirreps of $`๐^{\mathbf{(}๐\mathbf{)}}\mathbf{(}๐ฎ\mathbf{(}๐ต\mathbf{)}\mathbf{)}`$
As a consequence of Eq. (2.14), and of the assumption $`๐_2=I`$, the first Casimir operator $`U=๐_1`$ of $`๐^{(\lambda )}(G(N))`$ reduces to $`I`$; hence the eigenvalues of $`N`$ are integer. As usual, we shall restrict ourselves to those unirreps wherein they are nondegenerate (Rideau, 1992; Quesne and Vansteenkiste, 1996, 1997).<sup>2</sup>
Let us start with a normalized simultaneous eigenvector $`|c,n_0`$ of the Casimir operator $`๐=๐_3`$, defined in Eq. (2.12), and of the number operator $`N`$, corresponding to the eigenvalues $`c\text{}`$ and $`n_0\text{}`$, respectively. From Eqs. (2.5)โ(2.7), it results that as long as they are nonvanishing, the vectors
$$|c,n_0+n)=\{\begin{array}{cc}\left(a^{}\right)^n|c,n_0,\hfill & \text{if }n=0,1,\mathrm{}\text{,}\hfill \\ a^n|c,n_0,\hfill & \text{if }n=1,2,\mathrm{}\text{,}\hfill \end{array}$$
(3.1)
satisfy the relations
$$๐|c,n_0+n)=c|c,n_0+n),N|c,n_0+n)=(n_0+n)|c,n_0+n),$$
(3.2)
$$a^{}a|c,n_0+n)=\lambda _n|c,n_0+n),aa^{}|c,n_0+n)=\lambda _{n+1}|c,n_0+n),$$
(3.3)
where
$$\lambda _n=F(n_0+n)c.$$
(3.4)
In any unirrep, only nonnegative values of $`\lambda _n`$ are allowed. From Eq. (2.19), it is clear that the unirrep carrier space $`๐ฎ`$ is $`\text{}_\lambda `$-graded: $`๐ฎ=_{\mu =0}^{\lambda 1}๐ฎ_\mu `$, where $`๐ฎ_\mu =\{|c,n_0+n)n_0+n=\mu \mathrm{mod}\lambda \}`$. Hence, we have to discuss the unitarity conditions $`\lambda _n0`$ separately in each $`๐ฎ_\mu `$ subspace. Since the structure function $`F(N)`$ is an increasing linear function of $`N`$ in each $`๐ฎ_\mu `$, it is obvious that the algebra has no infinite-dimensional bounded from above (BFA) nor unbounded (UB) unirreps (Quesne and Vansteenkiste, 1996, 1997). It therefore only remains to successively consider the cases of infinite-dimensional bounded from below (BFB) unirreps and of finite-dimensional (FD) ones.
In the case of BFB unirreps, the eigenvalues of N are $`n_0`$, $`n_0+1`$, $`n_0+2`$, โฆ, and the unitarity conditions reduce to
$$\lambda _0=0,\lambda _n>0\mathrm{if}n=1,2,\mathrm{},\lambda 1.$$
(3.5)
The first condition in Eq. (3.5) fixes the Casimir operator eigenvalue,
$$c=n_0+\beta _{\mu _0},$$
(3.6)
where $`\mu _0\{0,1,\mathrm{},\lambda 1\}`$ is defined by
$$n_0=\mu _0\mathrm{mod}\lambda ,$$
(3.7)
while the second condition yields some restrictions on the algebra parameters,
$$\overline{\beta }_\nu \overline{\beta }_{\mu _0}+1>0,\mathrm{if}\nu =0,1,\mathrm{},\mu _01,$$
(3.8)
$$\overline{\beta }_\nu \overline{\beta }_{\mu _0}>0,\mathrm{if}\nu =\mu _0+1,\mu _0+2,\mathrm{},\lambda 1,$$
(3.9)
where
$$\overline{\beta }_\mu =\frac{\beta _\mu +\mu }{\lambda }.$$
(3.10)
In terms of the $`\alpha _\mu `$โs, Eqs. (3.6), (3.8), and (3.9) can be rewritten as
$$c=n_0+\underset{\nu =0}{\overset{\mu _01}{}}\alpha _\nu ,$$
(3.11)
and
$$\alpha _\nu <\lambda \mu _0+\nu \underset{\rho =\nu +1}{\overset{\mu _01}{}}\alpha _\rho ,\mathrm{if}\nu =0,1,\mathrm{},\mu _01,$$
(3.12)
$$\alpha _\nu >\mu _0\nu 1\underset{\rho =\mu _0}{\overset{\nu 1}{}}\alpha _\rho ,\mathrm{if}\nu =\mu _0,\mu _0+1,\mathrm{},\lambda 2,$$
(3.13)
respectively.
Whenever the unitarity conditions are satisfied, normalized basis states of $`๐ฎ`$ can be constructed from the vectors (3.1), and are given by
$$|c,n_0+n=\left[๐ฉ_n(c,n_0)\right]^{1/2}|c,n_0+n),n=0,1,2,\mathrm{},$$
(3.14)
where the normalization coefficient is
$$๐ฉ_n(c,n_0)=\underset{i=1}{\overset{n}{}}\lambda _i=\underset{i=1}{\overset{n}{}}[F(n_0+i)c].$$
(3.15)
By writing $`n`$ as $`n=k\lambda +\mu `$, where $`\mu \{0,1,\mathrm{},\lambda 1\}`$, and $`k\text{}`$, $`๐ฉ_n(c,n_0)`$ can be expressed in terms of gamma functions as
$`๐ฉ_{k\lambda +\mu }(c,n_0)`$ $`=`$ $`\lambda ^{k\lambda +\mu }\left({\displaystyle \underset{\nu =0}{\overset{\mu _0+\mu }{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+k+1)\right)\left({\displaystyle \underset{\nu =\mu _0+\mu +1}{\overset{\lambda 1}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+k)\right)`$ (3.16)
$`\times \left({\displaystyle \underset{\nu =0}{\overset{\mu _0}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+1)\right)^1\left({\displaystyle \underset{\nu =\mu _0+1}{\overset{\lambda 1}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0})\right)^1,`$
$`\mathrm{if}\mu =0,1,\mathrm{},\lambda \mu _01,`$
$`=`$ $`\lambda ^{k\lambda +\mu }\left({\displaystyle \underset{\nu =0}{\overset{\mu _0+\mu \lambda }{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+k+2)\right)\left({\displaystyle \underset{\nu =\mu _0+\mu \lambda +1}{\overset{\lambda 1}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+k+1)\right)`$
$`\times \left({\displaystyle \underset{\nu =0}{\overset{\mu _0}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0}+1)\right)^1\left({\displaystyle \underset{\nu =\mu _0+1}{\overset{\lambda 1}{}}}\mathrm{\Gamma }(\overline{\beta }_\nu \overline{\beta }_{\mu _0})\right)^1,`$
$`\mathrm{if}\mu =\lambda \mu _0,\lambda \mu _0+1,\mathrm{},\lambda 1.`$
In the case of FD unirreps, the eigenvalues of N are $`n_0`$, $`n_0+1`$, โฆ, $`n_0+d1`$, where the dimension $`d`$ may only take values in the set $`\{1,2,\mathrm{},\lambda 1\}`$. The unitarity conditions are then given by
$$\lambda _0=0,\lambda _n>0\mathrm{if}n=1,2,\mathrm{},d1,\lambda _d=0.$$
(3.17)
Defining $`\mu _0`$ and $`\overline{\beta }_\mu `$ as before by Eqs. (3.7) and (3.10), respectively, we obtain
$$c=n_0+\beta _{\mu _0},$$
(3.18)
$$\overline{\beta }_\nu \overline{\beta }_{\mu _0}>0,\mathrm{if}\nu =\mu _0+1,\mu _0+2,\mathrm{},\mu _0+d1,$$
(3.19)
$$\overline{\beta }_{\mu _0+d}\overline{\beta }_{\mu _0}=0,$$
(3.20)
for $`\mu _0=0`$, 1, โฆ, $`\lambda d1`$, and
$$c=n_0+d+\beta _{\mu _0\lambda +d},$$
(3.21)
$$\overline{\beta }_\nu \overline{\beta }_{\mu _0\lambda +d}>0,\mathrm{if}\nu =0,1,\mathrm{},\mu _0\lambda +d1,$$
(3.22)
$$\overline{\beta }_\nu \overline{\beta }_{\mu _0}>0,\mathrm{if}\nu =\mu _0+1,\mu _0+2,\mathrm{},\lambda 1,$$
(3.23)
$$\overline{\beta }_{\mu _0\lambda +d}\overline{\beta }_{\mu _0}+1=0,$$
(3.24)
for $`\mu _0=\lambda d`$, $`\lambda d+1`$, โฆ, $`\lambda 1`$. In terms of the algebra parameters $`\alpha _\mu `$, Eqs. (3.18)โ(3.20), and Eqs. (3.21)โ(3.24) become
$$c=n_0+\underset{\nu =0}{\overset{\mu _01}{}}\alpha _\nu ,$$
(3.25)
$$\alpha _\nu >\mu _0\nu 1\underset{\rho =\mu _0}{\overset{\nu 1}{}}\alpha _\rho ,\mathrm{if}\nu =\mu _0,\mu _0+1,\mathrm{},\mu _0+d2,$$
(3.26)
$$\alpha _{\mu _0+d1}=d\underset{\rho =\mu _0}{\overset{\mu _0+d2}{}}\alpha _\rho ,$$
(3.27)
for $`\mu _0=0`$, 1, โฆ, $`\lambda d1`$, and
$$c=n_0+d+\underset{\nu =0}{\overset{\mu _0\lambda +d1}{}}\alpha _\nu ,$$
(3.28)
$$\alpha _\nu <\lambda \mu _0+\nu d\underset{\rho =\nu +1}{\overset{\mu _0\lambda +d1}{}}\alpha _\rho ,\mathrm{if}\nu =0,1,\mathrm{},\mu _0\lambda +d1,$$
(3.29)
$$\alpha _\nu >\mu _0\nu 1\underset{\rho =\mu _0}{\overset{\nu 1}{}}\alpha _\rho ,\mathrm{if}\nu =\mu _0,\mu _0+1,\mathrm{},\lambda 2,$$
(3.30)
$$\alpha _{\mu _01}=d\underset{\rho =\mu _0\lambda +d}{\overset{\mu _02}{}}\alpha _\rho ,$$
(3.31)
for $`\mu _0=\lambda d`$, $`\lambda d+1`$, โฆ, $`\lambda 1`$, respectively.
Normalized basis states of the carrier space $`๐ฎ`$ of a $`d`$-dimensional unirrep are given by Eqs. (3.14) and (3.15), where $`n`$ now runs over the range $`n=0`$, 1, โฆ, $`d1`$. The corresponding normalization coefficient $`๐ฉ_n(c,n_0)`$ can be rewritten as
$$๐ฉ_n(c,n_0)=\lambda ^n\underset{\nu =\mu _0+1}{\overset{\mu _0+n}{}}\left(\overline{\beta }_\nu \overline{\beta }_{\mu _0}\right),$$
(3.32)
for $`\mu _0=0`$, 1, โฆ, $`\lambda d1`$, and
$`๐ฉ_n(c,n_0)`$ $`=`$ $`\lambda ^n{\displaystyle \underset{\nu =\mu _0+1}{\overset{\mu _0+n}{}}}\left(\overline{\beta }_\nu \overline{\beta }_{\mu _0\lambda +d}1\right),\mathrm{if}n=1,2,\mathrm{},\lambda \mu _01,`$ (3.33)
$`=`$ $`\lambda ^n\left({\displaystyle \underset{\nu =0}{\overset{\mu _0+n\lambda }{}}}\left(\overline{\beta }_\nu \overline{\beta }_{\mu _0\lambda +d}\right)\right)\left({\displaystyle \underset{\nu =\mu _0+1}{\overset{\lambda 1}{}}}\left(\overline{\beta }_\nu \overline{\beta }_{\mu _0\lambda +d}1\right)\right),`$
$`\mathrm{if}n=\lambda \mu _0,\lambda \mu _0+1,\mathrm{},d1,`$
for $`\mu _0=\lambda d`$, $`\lambda d+1`$, โฆ, $`\lambda 1`$.
In Tables I, II, and III, the detailed unirrep classification is given for $`\lambda =2`$, $`\lambda =3`$, and $`\lambda =4`$, respectively.
Of special interest in physical applications are the Fock-space unirreps, characterized by $`c=n_0=0`$. Since in this case $`\mu _0=0`$, such representations exist whenever the algebra parameters satisfy the conditions
$$\underset{\rho =0}{\overset{\nu }{}}\alpha _\rho >\nu 1,\mathrm{if}\nu =0,1,\mathrm{},\lambda 2,$$
(3.34)
in the BFB case, and
$$\underset{\rho =0}{\overset{\nu }{}}\alpha _\rho >\nu 1,\mathrm{if}\nu =0,1,\mathrm{},d2,$$
(3.35)
$$\underset{\rho =0}{\overset{d1}{}}\alpha _\rho =d,$$
(3.36)
in the FD one. The former are of bosonic type. Apart from the trivial one-dimensional unirrep, the latter are of fermionic or order-$`p`$-parafermionic type, according to whether $`d=2`$ or $`d=p+13`$. Note that parafermionic-type unirreps only appear for $`\lambda 4`$.
In the bosonic Fock-space representation, it may be interesting to consider a bosonic oscillator Hamiltonian (Quesne and Vansteenkiste, 1998), defined in appropriate units by
$$H_0=\frac{1}{2}\{a,a^{}\}.$$
(3.37)
By using Eqs. (2.5)โ(2.7), and (2.12), $`H_0`$ can be rewritten in the equivalent forms
$$H_0=a^{}a+\frac{1}{2}\left(I+\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu \right)=N+\frac{1}{2}I+\underset{\mu =0}{\overset{\lambda 1}{}}\gamma _\mu P_\mu ,$$
(3.38)
where the parameters $`\gamma _\mu `$ are defined by
$$\gamma _\mu \frac{1}{2}(\beta _\mu +\beta _{\mu +1})=\{\begin{array}{cc}\frac{1}{2}\alpha _0,\hfill & \text{if }\mu =0,\hfill \\ _{\nu =0}^{\mu 1}\alpha _\nu +\frac{1}{2}\alpha _\mu ,\hfill & \text{if }\mu =1,2,\mathrm{},\lambda 1.\hfill \end{array}$$
(3.39)
The latter satisfy the relation
$$\underset{\mu =0}{\overset{\lambda 1}{}}(1)^\mu \gamma _\mu =0,$$
(3.40)
deriving from Eq. (2.9), as well as the inequalities
$$\gamma _\mu >\frac{1}{2}(2\mu +1),\mathrm{if}\mu =0,1,\mathrm{},\lambda 2,$$
(3.41)
$$\gamma _{\lambda 1}>\frac{1}{2}(\lambda 1),$$
(3.42)
coming from conditions (3.34).
The states $`|n=|k\lambda +\mu `$, given by Eq. (3.14) where $`c=n_0=0`$, are the eigenstates of $`H_0`$, corresponding to the eigenvalues
$$E_{k\lambda +\mu }=k\lambda +\mu +\gamma _\mu +\frac{1}{2},k=0,1,2,\mathrm{},\mu =0,1,\mathrm{},\lambda 1.$$
(3.43)
In each $`_\mu =\{|k\lambda +\mu k=0,1,2,\mathrm{}\}`$ subspace of the $`\text{}_\lambda `$-graded Fock space $`=_{\mu =0}^{\lambda 1}_\mu `$, the spectrum of $`H_0`$ is harmonic, but the $`\lambda `$ infinite sets of equally spaced energy levels, corresponding to $`\mu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, may be shifted with respect to each other by some amounts depending upon the algebra parameters $`\alpha _0`$, $`\alpha _1`$, $`\mathrm{}`$$`\alpha _{\lambda 2}`$, through their linear combinations $`\gamma _0`$, $`\gamma _1`$, $`\mathrm{}`$$`\gamma _{\lambda 1}`$. As a result, one may get nondegenerate spectra, as well as spectra exhibiting some $`(\nu +1)`$-fold degeneracies, where $`\nu `$ may take any value in the set $`\{1,2,\mathrm{},\lambda 1\}`$ (Quesne and Vansteenkiste, 1998, 1999).
### 3.2 Unirreps of $`๐_{๐ถ_\mathrm{๐}๐ถ_\mathrm{๐}\mathbf{}๐ถ_{๐\mathbf{}\mathrm{๐}}}^{\mathbf{(}๐\mathbf{)}}`$
Let us now turn ourselves to the general case of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, defined in Eqs. (2.1) and (2.2). Since we do not assume $`๐_2=I`$, the eigenvalues of $`N`$ are not restricted to integer values anymore. It can however be shown that they are discrete. The proof proceeds as in Jordan et al. (1963) and Quesne and Vansteenkiste (1997), and can be summarized as follows. The Casimir operator $`๐_1`$, defined in Eq. (2.10), is unitary, so that in any given unirrep its eigenvalue can be written as $`\mathrm{exp}(\mathrm{i2}\pi \nu _0)`$, where $`\nu _0\text{}`$. On the other hand, the eigenvalues of $`๐_1`$ can be determined from those of the Hermitian operator $`N`$. The spectral mapping theorem leads to eigenvalues of $`๐_1`$ of the form $`\mathrm{exp}(\mathrm{i2}\pi x)`$, where $`x\text{}`$ are the eigenvalues of $`N`$. The equivalence of the two expressions for the eigenvalues of $`๐_1`$ implies that $`x=\nu _0+n`$, $`n\text{}`$, in any given unirrep, which completes the proof. As in Section 3.1, we shall restrict ourselves to those unirreps wherein the spectrum of $`N`$ is not only discrete, but also nondegenerate.
As in Eq. (3.1), the carrier space of any $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ unirrep can be constructed by successive applications of $`a^{}`$ or $`a`$ on a normalized simultaneous eigenvector $`|c,\gamma ,\nu _0`$ of $`N`$ and of the Casimir operators $`๐_1`$, $`๐_2`$, $`๐_3`$, defined in Eqs. (2.10)โ(2.12),
$$N|c,\gamma ,\nu _0=\nu _0|c,\gamma ,\nu _0,$$
(3.44)
$$๐_1|c,\gamma ,\nu _0=e^{\mathrm{i2}\pi r_0}|c,\gamma ,\nu _0,$$
(3.45)
$$๐_2|c,\gamma ,\nu _0=e^{\mathrm{i2}\pi (r_0+\gamma )/\lambda }|c,\gamma ,\nu _0,$$
(3.46)
$$๐_3|c,\gamma ,\nu _0=c|c,\gamma ,\nu _0.$$
(3.47)
Here $`c`$, $`\nu _0\text{}`$, $`\gamma \{0,1,\mathrm{},\lambda 1\}`$, $`r_0[0,1)`$ is defined by
$$\nu _0=n_0+r_0,n_0\text{},$$
(3.48)
and the eigenvalue of $`๐_2`$ is determined from Eq. (2.14).
Let us now introduce some new operators and parameters, defined by
$$N^{}Nr_0I,a^{}a^{},a^{}a,T^{}e^{\mathrm{i2}\pi \gamma /\lambda }T,$$
(3.49)
$$\kappa _\mu ^{}e^{\mathrm{i2}\pi \mu \gamma /\lambda }\kappa _\mu =\kappa _{\lambda \mu }^{},$$
(3.50)
from which we obtain
$$P_\mu ^{}\frac{1}{\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu \nu /\lambda }T^\nu =P_{\mu +\gamma },$$
(3.51)
$$\alpha _\mu ^{}\underset{\nu =1}{\overset{\lambda 1}{}}e^{\mathrm{i2}\pi \mu \nu /\lambda }\kappa _\nu ^{}=\alpha _{\mu +\gamma }=\alpha _\mu ^{}.$$
(3.52)
It is obvious that $`N^{}`$, $`a^{}`$, $`a^{}`$, $`T^{}`$ (or $`P_\mu ^{}`$) satisfy the defining relations (2.1) and (2.2) (or (2.5)โ(2.7)) of $`๐_{\alpha _0^{}\alpha _1^{}\mathrm{}\alpha _{\lambda 2}^{}}^{(\lambda )}`$, where the primed parameters $`\alpha _\mu ^{}`$ are given by Eq. (3.52). The corresponding Casimir operators $`๐_1^{}`$, $`๐_2^{}`$, $`๐_3^{}`$ are found to be expressible in terms of the old ones $`๐_1`$, $`๐_2`$, $`๐_3`$,
$$๐_1^{}e^{\mathrm{i2}\pi N^{}}=e^{\mathrm{i2}\pi r_0}๐_1,$$
(3.53)
$$๐_2^{}e^{\mathrm{i2}\pi N^{}/\lambda }T^{}=e^{\mathrm{i2}\pi (r_0\gamma )/\lambda }๐_2,$$
(3.54)
$$๐_3^{}N^{}+\underset{\mu =0}{\overset{\lambda 1}{}}\beta _\mu ^{}P_\mu ^{}a^{}a^{}=๐_3(r_0+\beta _\gamma )I,$$
(3.55)
where $`\beta _\mu ^{}_{\nu =0}^{\mu 1}\alpha _\nu ^{}=\beta _{\mu +\gamma }\beta _\gamma `$.
Hence, the simultaneous eigenvector $`|c,\gamma ,\nu _0`$ of $`N`$, $`๐_1`$, $`๐_2`$, $`๐_3`$ is also a simultaneous eigenvector of $`N^{}`$, $`๐_1^{}`$, $`๐_2^{}`$, $`๐_3^{}`$, satisfying the relations
$$N^{}|c,\gamma ,\nu _0=n_0|c,\gamma ,\nu _0,$$
(3.56)
$$๐_1^{}|c,\gamma ,\nu _0=๐_2^{}|c,\gamma ,\nu _0=|c,\gamma ,\nu _0,$$
(3.57)
$$๐_3^{}|c,\gamma ,\nu _0=c^{}|c,\gamma ,\nu _0,$$
(3.58)
where
$$c^{}=cr_0\beta _\gamma .$$
(3.59)
From Section 3.1, it follows that such a state may be identified with the starting eigenvector $`|c^{},n_0`$ of some unirrep of the GDOA $`๐^{(\lambda )}(G^{}(N^{}))`$, where $`G^{}(N^{})=I+_{\mu =0}^{\lambda 1}\alpha _\mu ^{}P_\mu ^{}`$. Since $`a^{}=a^{}`$ and $`a^{}=a`$, this correspondence between $`|c,\gamma ,\nu _0`$ and $`|c^{},n_0`$ extends to the remaining basis states of the $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ and $`๐^{(\lambda )}(G^{}(N^{}))`$ unirreps built on such vectors, respectively.
We conclude that to every BFB (or FD) unirrep of $`๐^{(\lambda )}(G^{}(N^{}))`$, specified by some minimal $`N^{}`$ eigenvalue $`n_0\text{}`$ (and some dimension $`d`$), we may associate an infinite number of BFB (or FD) unirreps of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, characterized by minimal $`N`$ eigenvalues $`\nu _0=n_0+r_0`$, $`r_0[0,1)`$, as well as $`๐_2`$ eigenvalues $`\mathrm{exp}[\mathrm{i2}\pi (r_0+\gamma )/\lambda ]`$, $`\gamma \{0,1,\mathrm{},\lambda 1\}`$ (and the same dimension $`d`$). The eigenvalues of the corresponding Casimir operators $`๐_3^{}=๐^{}`$ and $`๐_3`$ are connected by Eq. (3.59). Furthermore, all the $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ unirreps are obtained by this mapping procedure.
## 4 APPLICATION OF $`๐ช_๐`$-EXTENDED OSCILLATOR ALGEBRAS TO PSSQM OF ORDER $`๐\mathbf{=}๐\mathbf{}\mathrm{๐}`$
PSSQM of order two was introduced by Rubakov and Spiridonov (1988) as a generalization of SSQM (Witten, 1981), obtained by combining standard fermions with parafermions of order two (Green, 1953; Ohnuki and Kamefuchi, 1982) instead of standard fermions. Its extension to arbitrary order $`p`$, due to Khare (1992, 1993), is described in terms of parasupercharge operators $`Q`$, $`Q^{}`$, and a parasupersymmetric Hamiltonian $``$, satisfying the relations
$$Q^{p+1}=0(\mathrm{with}Q^p0),$$
(4.1)
$$[,Q]=0,$$
(4.2)
$$Q^pQ^{}+Q^{p1}Q^{}Q+\mathrm{}+QQ^{}Q^{p1}+Q^{}Q^p=2pQ^{p1},$$
(4.3)
and their Hermitian conjugates.
As shown by Bagchi et al. (1997), PSSQM of order $`p`$ can be reformulated in terms of $`p`$ super (rather than parasuper) charges $`Q_\nu `$, $`\nu =1`$, 2, โฆ, $`p`$, all of which satisfy $`Q_\nu ^2=0`$ and commute with $``$. However, unlike in usual SSQM, $``$ cannot be simply expressed in terms of the $`p`$ supercharges (except in a very special case to be reviewed below). More specifically, let us set
$$Q=\underset{\nu =1}{\overset{p}{}}\sigma _\nu Q_\nu ,$$
(4.4)
where $`\sigma _\nu `$ are some complex constants, and $`Q_\nu `$, $`\nu =1`$, 2, โฆ, $`p`$, are assumed to satisfy the relations
$$Q_\nu Q_\nu ^{}=\delta _{\nu ^{},\nu +1}Q_\nu Q_{\nu +1},$$
(4.5)
$$Q_\nu Q_\nu ^{}^{}=\delta _{\nu ^{},\nu }Q_\nu Q_\nu ^{},$$
(4.6)
$$Q_\nu ^{}Q_\nu ^{}=\delta _{\nu ^{},\nu }Q_\nu ^{}Q_\nu .$$
(4.7)
Then, the operator $`Q`$, defined in Eq. (4.4), satisfies Eqs. (4.1)โ(4.3) if
$$\sigma _\nu 0,\nu =1,2,\mathrm{},p,$$
(4.8)
$$[,Q_\nu ]=0,\nu =1,2,\mathrm{},p,$$
(4.9)
$$\left(\underset{\nu =1}{\overset{p1}{}}\sigma _\nu \right)Q_1\mathrm{\Sigma }+\left(\underset{\nu =2}{\overset{p}{}}\sigma _\nu \right)\mathrm{\Sigma }Q_p=2p\left[\left(\underset{\nu =1}{\overset{p1}{}}\sigma _\nu \right)Q_1Q_2\mathrm{}Q_{p1}+\left(\underset{\nu =2}{\overset{p}{}}\sigma _\nu \right)Q_2Q_3\mathrm{}Q_p\right],$$
(4.10)
where
$$\mathrm{\Sigma }|\sigma _1|^2Q_1^{}Q_1Q_2\mathrm{}Q_{p1}+\underset{\nu =2}{\overset{p1}{}}|\sigma _\nu |^2Q_2Q_3\mathrm{}Q_\nu Q_\nu ^{}Q_\nu Q_{\nu +1}\mathrm{}Q_{p1}+|\sigma _p|^2Q_2Q_3\mathrm{}Q_pQ_p^{}.$$
(4.11)
In the standard realization of PSSQM related to parafermions of order $`p`$ (Khare, 1992, 1993), $`\sigma _\nu =1`$, and $`Q_\nu `$, $`Q_\nu ^{}`$, $``$ are represented by $`(p+1)\times (p+1)`$ matrices, whose elements are
$$(Q_\nu )_{\alpha ,\beta }=(P\mathrm{i}W_\beta )\delta _{\alpha ,\beta +1}\delta _{\beta ,p+1\nu },$$
(4.12)
$$\left(Q_\nu ^{}\right)_{\alpha ,\beta }=(P+\mathrm{i}W_\alpha )\delta _{\alpha ,p+1\nu }\delta _{\beta ,\alpha +1},$$
(4.13)
$$()_{\alpha ,\beta }=_\alpha \delta _{\alpha ,\beta },$$
(4.14)
where $`\alpha `$, $`\beta =1`$, 2, โฆ, $`p+1`$. Here $`P=\mathrm{i}/x`$ is the momentum operator, $`W_\nu (x)`$, $`\nu =1`$, 2, โฆ, $`p`$, are superpotentials, and
$$_\nu =\frac{1}{2}\left(P^2+W_\nu ^2W_\nu ^{}+C_\nu \right),\nu =1,2,\mathrm{},p,$$
(4.15)
$$_{p+1}=\frac{1}{2}\left(P^2+W_p^2+W_p^{}+C_p\right),$$
(4.16)
with $`C_\nu \text{}`$. The operator-valued matrices (4.12) and (4.13) automatically satisfy Eqs. (4.5)โ(4.7), while Eqs. (4.9) and (4.10) impose the conditions
$$W_\nu ^2+W_\nu ^{}+C_\nu =W_{\nu +1}^2W_{\nu +1}^{}+C_{\nu +1},\nu =1,2,\mathrm{},p1,$$
(4.17)
and
$$\underset{\nu =1}{\overset{p}{}}C_\nu =0,$$
(4.18)
respectively.
For arbitrary $`W_\nu `$โs satisfying Eqs. (4.17) and (4.18), the spectrum of the parasupersymmetric Hamiltonian $``$ is ($`p+1`$)-fold degenerate at least starting from the $`p`$th excited state onwards. The nature of the ground and the first ($`p1`$) excited states however depends on the specific form of the $`W_\nu `$โs. For the special choice $`W_1=W_2=\mathrm{}=W_p=\omega x`$, $``$ becomes the parasupersymmetric oscillator Hamiltonian, which can be realized in terms of bosons and parafermions of order $`p`$. Its ground state is nondegenerate, and has a negative energy, while the $`\nu `$th excited state for $`\nu =1`$, 2, โฆ, $`p1`$, is ($`\nu +1`$)-fold degenerate.
We now plan to show that the PSSQM algebra (4.1)โ(4.3) can be realized in terms of the generators of $`๐^{(\lambda )}(G(N))`$, $`\lambda =p+1`$, in their bosonic Fock-space representation (thence the parameters $`\alpha _0`$, $`\alpha _1`$, โฆ, $`\alpha _{\lambda 2}`$ satisfy Eq. (3.34)). This will prove that PSSQM of arbitrary order $`p`$ can be bosonized, as is the case for standard SSQM (Brzeziลski et al., 1993; Plyushchay, 1996a, b; Beckers et al., 1997), and PSSQM of order two (Quesne and Vansteenkiste, 1998).
In view of the results previously obtained for $`p=2`$ (Quesne and Vansteenkiste, 1998), let us take as ansรคtze for the operators $`Q`$ and $``$ the expressions
$$Q=\underset{\nu =1}{\overset{p}{}}\eta _{\mu +\nu }a^{}P_{\mu +\nu },$$
(4.19)
$$=H_0+\frac{1}{2}\underset{\nu =0}{\overset{p}{}}r_\nu P_\nu ,$$
(4.20)
where $`H_0`$ is the bosonic oscillator Hamiltonian (3.37) associated with the algebra $`๐^{(p+1)}(G(N))`$, $`\eta _{\mu +\nu }`$, $`\nu =1`$, 2, โฆ, $`p`$, are some complex constants, and $`r_\nu `$, $`\nu =0`$, 1, โฆ, $`p`$, some real ones. The purpose of the last term on the right-hand side of Eq. (4.20) is to make the $`p+1`$ families of $`H_0`$ equally spaced eigenvalues coincide at least starting from the $`p`$th excited state onwards. Note that in Eqs. (4.19) and (4.20), $`\mu `$ takes some fixed, arbitrary value in the set $`\{0,1,\mathrm{},p\}`$. The operators $`Q`$, $`Q^{}`$, $``$, and all the quantities to be considered hereafter, depend on this $`\mu `$ value, although for simplicityโs sake we chose not to explicitly exhibit such a dependence by appending a $`\mu `$ index to them.
It is straightforward to see that the operators
$$Q_\nu =a^{}P_{p+1+\mu \nu },\nu =1,2,\mathrm{},p,$$
(4.21)
satisfy Eqs. (4.5)โ(4.7); hence $`Q`$, as defined by Eq. (4.19), can be written in the form (4.4) by setting
$$\sigma _\nu =\eta _{p+1+\mu \nu },\nu =1,2,\mathrm{},p.$$
(4.22)
Equation (4.8) leads to the restriction
$$\eta _{\mu +\nu }0,\nu =1,2,\mathrm{},p.$$
(4.23)
After some calculations, one finds that Eqs. (4.9) and (4.10) are equivalent to the conditions
$$r_{\mu +\nu }=2+\alpha _{\mu +\nu }+\alpha _{\mu +\nu +1}+r_{\mu +\nu +1},\nu =1,2,\mathrm{},p,$$
(4.24)
and
$$\underset{\nu =1}{\overset{p}{}}|\eta _{\mu +\nu }|^2=2p,$$
(4.25)
$$\underset{\nu =2}{\overset{p}{}}|\eta _{\mu +\nu }|^2\left(\nu 1+\underset{\rho =0}{\overset{\nu 2}{}}\alpha _{\mu +\rho +2}\right)=p(1+\alpha _{\mu +2}+r_{\mu +2}),$$
(4.26)
respectively.
Equation (4.24) is a nonhomogeneous system of $`p`$ linear equations in ($`p+1`$) unknowns $`r_{\mu +\nu }`$, $`\nu =0`$, 1, โฆ, $`p`$. Its solution yields $`p`$ of them in terms of the remaining one, e.g., $`r_\mu `$, $`r_{\mu +1}`$, $`r_{\mu +3}`$, โฆ, $`r_{\mu +p}`$ in terms of $`r_{\mu +2}`$:
$`r_\mu `$ $`=`$ $`2(p1)\alpha _\mu \alpha _{\mu +2}2{\displaystyle \underset{\rho =3}{\overset{p}{}}}\alpha _{\mu +\rho }+r_{\mu +2}`$ (4.27)
$`=`$ $`2(p1)2\gamma _\mu +2\gamma _{\mu +2}+r_{\mu +2},`$
$`r_{\mu +1}`$ $`=`$ $`2+\alpha _{\mu +1}+\alpha _{\mu +2}+r_{\mu +2}=22\gamma _{\mu +1}+2\gamma _{\mu +2}+r_{\mu +2},`$ (4.28)
$`r_{\mu +\nu }`$ $`=`$ $`2(\nu 2)\alpha _{\mu +2}2{\displaystyle \underset{\rho =3}{\overset{\nu 1}{}}}\alpha _{\mu +\rho }\alpha _{\mu +\nu }+r_{\mu +2}`$ (4.29)
$`=`$ $`2(\nu 2)+2\gamma _{\mu +2}2\gamma _{\mu +\nu }+r_{\mu +2},\nu =3,4,\mathrm{},p,`$
where $`\gamma _\mu `$ is defined in Eq. (3.39).
Equation (4.25) restricts the range of $`|\eta _{\mu +\nu }|^2`$, $`\nu =1`$, 2, โฆ, $`p`$, while Eq. (4.26) fixes the value of $`r_{\mu +2}`$ in terms of the latter and the algebra parameters. We conclude that it is possible to find values of $`\eta _{\mu +\nu }`$ and $`r_\nu `$ in Eqs. (4.19) and (4.20), so that Eqs. (4.1)โ(4.3) are satisfied. Choosing for instance
$$|\eta _{\mu +\nu }|^2=2,\nu =1,2,\mathrm{},p,$$
(4.30)
we obtain
$$r_{\mu +2}=\frac{1}{p}\left[(p2)\alpha _{\mu +2}+2\underset{\nu =3}{\overset{p}{}}(p\nu +1)\alpha _{\mu +\nu }+p(p2)\right],$$
(4.31)
or
$`r_{\mu +2}`$ $`=`$ $`{\displaystyle \frac{1}{p}}\{2[1(1)^p]{\displaystyle \underset{\nu =0}{\overset{\mu +1}{}}}(1)^{\mu +1\nu }\gamma _\nu 2[p1(1)^p]\gamma _{\mu +2}`$ (4.32)
$`+2{\displaystyle \underset{\nu =3}{\overset{p2}{}}}[1+(1)^{p\nu }]\gamma _{\mu +\nu }+4\gamma _{\mu +p}+p(p2)\}.`$
In going from Eq. (4.31) to Eq. (4.32), we used the inverse of Eq. (3.39), namely
$$\alpha _\mu =\{\begin{array}{cc}2\gamma _0,\hfill & \text{if }\mu =0,\hfill \\ 4_{\nu =0}^{\mu 1}(1)^{\mu \nu }\gamma _\nu +2\gamma _\mu ,\hfill & \text{if }\mu =1,2,\mathrm{},\lambda 1.\hfill \end{array}$$
(4.33)
From Eqs. (3.38), and (4.27)โ(4.29), it follows that the parasupersymmetric Hamiltonian (4.20) can be rewritten as
$$=N+\frac{1}{2}(2\gamma _{\mu +2}+r_{\mu +2}2p+3)I+\underset{\nu =1}{\overset{p}{}}(p+1\nu )P_{\mu +\nu },$$
(4.34)
where $`r_{\mu +2}`$ is given by Eq. (4.32). The eigenstates $`|n=|k(p+1)+\nu `$, $`n`$, $`k=0`$, 1, 2, โฆ, $`\nu =0`$, 1, โฆ, $`p`$, of $`H_0`$ are also eigenstates of $``$, corresponding to the eigenvalues
$$_{k(p+1)+\nu }=k(p+1)+\frac{1}{2}(2\gamma _{\mu +2}+r_{\mu +2}+2\mu 2p+3),\mathrm{if}\nu =0,1,\mathrm{},\mu ,$$
(4.35)
$$_{k(p+1)+\nu }=(k+1)(p+1)+\frac{1}{2}(2\gamma _{\mu +2}+r_{\mu +2}+2\mu 2p+3),\mathrm{if}\nu =\mu +1,\mu +2,\mathrm{},p.$$
(4.36)
All the levels are therefore equally spaced. The ground state, corresponding to the energy
$$_0=_1=\mathrm{}=_\mu =\frac{1}{2}(2\gamma _{\mu +2}+r_{\mu +2}+2\mu 2p+3),$$
(4.37)
is ($`\mu +1`$)-fold degenerate, whereas the excited states are ($`p+1`$)-fold degenerate. Note that since $`\mu `$ may take any value in the set $`\{0,1,\mathrm{},p\}`$, the ground-state degeneracy may accordingly vary between 1 and $`p+1`$. Unbroken (resp. broken) PSSQM corresponds to $`\mu =0`$ (resp. $`\mu =1`$, 2, โฆ, or $`p`$).
To study the sign of the ground-state energy, we have to insert Eq. (4.32) into Eq. (4.37). The result reads
$`_0`$ $`=`$ $`_1=\mathrm{}=_\mu ={\displaystyle \frac{1}{2p}}\left[4{\displaystyle \underset{\nu =0}{\overset{(\mu 2)/2}{}}}\gamma _{2\nu +1}+4{\displaystyle \underset{\nu =(\mu +2)/2}{\overset{[p/2]}{}}}\gamma _{2\nu }+p(2\mu p+1)\right],`$ (4.38)
$`\mathrm{if}\mu =0,2,\mathrm{},2[p/2],`$
$`_0`$ $`=`$ $`_1=\mathrm{}=_\mu ={\displaystyle \frac{1}{2p}}\left[4{\displaystyle \underset{\nu =0}{\overset{(\mu 1)/2}{}}}\gamma _{2\nu }+4{\displaystyle \underset{\nu =(\mu +1)/2}{\overset{[(p1)/2]}{}}}\gamma _{2\nu +1}+p(2\mu p+1)\right],`$ (4.39)
$`\mathrm{if}\mu =1,3,\mathrm{},2[(p1)/2]+1,`$
where $`[a]`$ denotes the largest integer contained in $`a`$, and $`_{\nu =a}^b0`$ if $`a>b`$. From the conditions (3.41) and (3.42) for the existence of the bosonic Fock-space representation, it follows that
$$_0=_1=\mathrm{}=_\mu >\frac{1}{p}(p+1)(\mu p+1),\mathrm{if}\mu =0,1,\mathrm{},p2,$$
(4.40)
$$_0=_1=\mathrm{}=_\mu >0,\mathrm{if}\mu =p1,p.$$
(4.41)
Since the right-hand side of Eq. (4.40) is negative, for $`\mu =0`$, 1, โฆ, $`p2`$, the ground-state energy may be positive, null, or negative according to the values taken by the algebra parameters. We therefore recover a well-known property of PSSQM of order $`p2`$: unlike in SSQM (corresponding to $`p=1`$), the energy eigenvalues are not necessarily nonnegative, and there is no connection between the nonvanishing (resp. vanishing) ground-state energy and the broken (resp. unbroken) PSSQM.
As noted by Khare et al. (1993b), there is however a special case in the standard PSSQM realization (4.12)โ(4.16), wherein this unsatisfactory situation does not occur, and moreover the parasupersymmetric Hamiltonian $``$ can be expressed directly in terms of the parasupercharge operators $`Q`$, $`Q^{}`$, in contrast with Eq. (4.3). Whenever, in Eq. (4.17), all the constants $`C_\nu `$ vanish, one can indeed write $``$ as
$$=\frac{1}{2}\left[\left(Q^{}QQQ^{}\right)^2+Q^{}Q^2Q^{}\right]^{1/2},$$
(4.42)
whose eigenvalues are necessarily nonnegative. Furthermore, its ground-state energy vanishes (resp. is positive) for unbroken (resp. broken) PSSQM.
Such a special case does have a counterpart in the present bosonic realization. By introducing Eqs. (4.19) and (4.20) into Eq. (4.42), and taking Eq. (4.30) into account, it is easy to show that Eq. (4.42) is equivalent to the following additional conditions,
$$r_\mu =1\alpha _\mu ,r_{\mu +1}=1+\alpha _{\mu +1},r_{\mu +\nu }=0,\nu =2,3,\mathrm{},p,$$
(4.43)
$$\alpha _{\mu +\nu }=1,\nu =2,3,\mathrm{},p,$$
(4.44)
which can be checked to be compatible with the previous ones, given in Eqs. (4.27)โ(4.29), and (4.31).
However, the conditions (3.34) for the existence of the bosonic Fock-space representation are compatible with Eq. (4.44) only for $`\mu =0`$ and $`\mu =p`$. In the former case, $`\alpha _1=p1\alpha _0`$, $`\alpha _2=\alpha _3=\mathrm{}=\alpha _p=1`$, where $`\alpha _0>1`$, and from Eqs. (4.34) and (4.43),
$$=N+\underset{\nu =1}{\overset{p}{}}(p+1\nu )P_\nu .$$
(4.45)
PSSQM is then unbroken, and the ground-state energy vanishes ($`_0=0`$ in accordance with Eq. (4.37), since $`\gamma _2=p\frac{3}{2}`$). In the latter case, $`\alpha _1=\alpha _2=\mathrm{}=\alpha _{p1}=1`$, $`\alpha _p=p1\alpha _0`$, where $`\alpha _0>1`$, and
$$=N+\underset{\nu =0}{\overset{p}{}}(\alpha _0+1\nu )P_\nu .$$
(4.46)
PSSQM is then broken, and the ground-state energy $`_0=_1=\mathrm{}=_p=\alpha _0+1`$ (in accordance with Eq. (4.37), since $`\gamma _{p+2}=\gamma _1=\alpha _0\frac{1}{2}`$) is positive, the ground state being ($`p+1`$)-fold degenerate as all the excited states.
Furthermore, by using conditions (4.43) and (4.44), it can be shown that $``$ can be rewritten in terms of the supercharges (4.21) as
$$=Q_1Q_1^{}+\underset{\nu =1}{\overset{p}{}}Q_\nu ^{}Q_\nu .$$
(4.47)
This result has also its counterpart in the standard PSSQM realization (Bagchi et al., 1997).
Going back now to the general case corresponding to conditions (4.27)โ(4.31) only, we note that Eq. (4.30), yielding the coefficients in the expansion of the parasupercharges (4.19), has many solutions. This is not surprising since Khare did show that in the standard PSSQM realization (4.12)โ(4.16), $``$ has in fact $`p`$ (and not only one) conserved parasupercharges, as well as $`p`$ bosonic constants (Khare, 1992, 1993). In other words, there exist $`p`$ independent operators $`Q_r`$, $`r=1`$, 2, โฆ, $`p`$, satisfying with $``$ the set of equations (4.1)โ(4.3), and $`p`$ other independent operators $`I_t`$, $`t=2`$, 3, โฆ, $`p+1`$, commuting with $``$, as well as among themselves. The former are obtained from Eqs. (4.4) and (4.12) by setting $`\sigma _\nu =1`$ for $`r=1`$, and $`\sigma _\nu =12\delta _{\nu ,p+1r}`$ for $`r=2`$, 3, โฆ, $`p`$, while the latter are given by $`(I_t)_{\alpha ,\beta }=\delta _{\alpha ,\beta }(12\delta _{\alpha ,t})`$, where $`t=2`$, 3, โฆ, $`p+1`$, and $`\alpha `$, $`\beta =1`$, 2, โฆ, $`p+1`$. In addition, for any $`r_k`$, $`r_{k+1}`$, โฆ, $`r_{k+p}\{1,2,\mathrm{},p\}`$,
$$Q_{r_k}Q_{r_{k+1}}\mathrm{}Q_{r_{k+p}}=0,$$
(4.48)
and for any $`r\{1,2,\mathrm{},p\}`$, $`t\{2,3,\mathrm{},p+1\}`$,
$$[I_t,Q_r]=\underset{s=1}{\overset{p}{}}d_{tr}^sQ_s,$$
(4.49)
where $`d_{tr}^s`$ are some real constants, e.g.,
$$d_{21}^1=d_{22}^2=0,d_{21}^2=d_{22}^1=2,d_{31}^1=d_{31}^2=d_{32}^1=d_{32}^2=1,$$
(4.50)
for $`p=2`$. Finally, the $`Q_r`$โs satisfy some mixed multilinear relations generalizing Eq. (4.3), and involving $``$ and the bosonic constants $`I_t`$. For $`p=2`$, for instance, there are six such independent relations
$$I_3Q_s^2Q_r^{}+Q_sQ_r^{}Q_s+I_2Q_r^{}Q_s^2=4Q_r,$$
(4.51)
$$Q_rQ_sQ_s^{}+Q_sQ_s^{}Q_r+I_2Q_s^{}Q_rQ_s=4Q_r,$$
(4.52)
$$I_3Q_sQ_rQ_s^{}+Q_rQ_s^{}Q_s+Q_s^{}Q_sQ_r=4Q_r,$$
(4.53)
where $`(r,s)=(1,2)`$, $`(2,1)`$.
It is straightforward to show that the operators $`Q_r`$ and $`I_t`$ have also their counterpart in the present bosonic realization. Let us indeed consider the operators
$$Q_r=\sqrt{2}\underset{\nu =1}{\overset{p}{}}b_r^\nu a^{}P_{\mu +\nu },r=1,2,\mathrm{},p,$$
(4.54)
$$I_t=\underset{\nu =1}{\overset{p+1}{}}b_t^\nu P_{\mu +\nu },t=1,2,\mathrm{},p+1,$$
(4.55)
where
$$b_t^\nu =12\delta _{t,\nu }(1\delta _{t,1}),t,\nu =1,2,\mathrm{},p+1.$$
(4.56)
The $`b_t^\nu `$โs taking values only in the set $`\{1,+1\}`$, it is clear that each $`Q_r`$ in Eq. (4.54) satisfies the PSSQM algebra (4.1)โ(4.3) with Hamiltonian (4.34). It is also obvious that $`I_2`$, $`I_3`$, โฆ, $`I_{p+1}`$, as defined by Eq. (4.55), commute with the same, as well as among themselves, while $`I_1`$ reduces to the unit operator. Equation (4.48) directly follows for $`n=p`$ from the relation
$$Q_{r_k}Q_{r_{k+1}}\mathrm{}Q_{r_{k+n}}=2^{(n+1)/2}\left(a^{}\right)^{n+1}\underset{\nu =1}{\overset{pn}{}}B_\nu (r_k,r_{k+1},\mathrm{},r_{k+n})P_{\mu +\nu },$$
(4.57)
$$B_\nu (r_k,r_{k+1},\mathrm{},r_{k+n})\underset{l=0}{\overset{n}{}}b_{r_{k+l}}^{\nu +nl},$$
(4.58)
which can be proved by induction over $`n`$.
Considering now Eq. (4.49), we obtain from Eqs. (4.54) and (4.55)
$$[I_t,Q_r]=\sqrt{2}a^{}\underset{\nu =1}{\overset{p}{}}c_{tr}^\nu P_{\mu +\nu },c_{tr}^\nu \left(b_t^{\nu +1}b_t^\nu \right)b_r^\nu .$$
(4.59)
By combining this result with the inverse of Eq. (4.54),
$$\sqrt{2}a^{}P_{\mu +\nu }=\underset{r=1}{\overset{p}{}}b_\nu ^rQ_r,$$
(4.60)
$$b_\nu ^r\frac{1}{2}\{\delta _{\nu ,1}[1+(2p)\delta _{r,1}]+(1\delta _{\nu ,1})(\delta _{r,1}\delta _{\nu ,r})\},$$
(4.61)
we get Eq. (4.49) with $`d_{tr}^s`$ given by
$$d_{tr}^s=\underset{\nu =1}{\overset{p}{}}c_{tr}^\nu b_\nu ^s.$$
(4.62)
For the special cases $`p=2`$ and $`p=3`$, considered by Khare (1992, 1993), this general formula yields the correct results (see e.g. Eq. (4.50)).
Finally, for the mixed multilinear relations satisfied by the $`Q_r`$โs and $`I_t`$โs, let us consider a general relation of the type
$`I_{t_1}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_p}Q_s^{}+I_{t_2}Q_{r_2}Q_{r_3}\mathrm{}Q_{r_p}Q_s^{}Q_{r_1}+\mathrm{}+I_{t_p}Q_{r_p}Q_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_{p1}}`$
$`+I_{t_{p+1}}Q_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_p}=2pQ_r^{p1},`$ (4.63)
where $`r_1`$, $`r_2`$, โฆ, $`r_p\{1,2,\mathrm{},p\}`$, and $`t_1`$, $`t_2`$, โฆ, $`t_{p+1}\{1,2,\mathrm{},p+1\}`$. It is clear that such a relation cannot be valid for any choice of the indices in the ranges indicated. To find to which choices it applies when definitions (4.54) and (4.55) are used, let us work out the conditions implied by Eq. (4.63).
After some calculations, one gets
$$\underset{\nu =1}{\overset{p}{}}D_k^\nu =pB_k\left([r]^{p1}\right),k=1,2,$$
(4.64)
$$\underset{\nu =2}{\overset{p}{}}D_k^\nu \left(\nu 1+\underset{\rho =0}{\overset{\nu 2}{}}\alpha _{\mu +\rho +2}\right)=\frac{p}{2}B_k\left([r]^{p1}\right)(1+\alpha _{\mu +2}+r_{\mu +2}),k=1,2,$$
(4.65)
where $`[r]^{p1}`$ means that $`r`$ is repeated ($`p1`$) times, and
$$D_k^\nu b_{t_{\nu +2k}}^{p+k1}B_\nu (r_{\nu +2k},r_{\nu +3k},\mathrm{},r_p)b_s^\nu B_k(r_1,r_2,\mathrm{},r_{\nu +1k}).$$
(4.66)
Since $`B_k\left([r]^{p1}\right)`$ and $`D_k^\nu `$ take values in the set $`\{+1,1\}`$, Eq. (4.64) is satisfied if and only if
$$D_k^\nu =B_k\left([r]^{p1}\right),k=1,2,\nu =1,2,\mathrm{},p.$$
(4.67)
Then Eq. (4.65) reduces to Eq. (4.26), where the choice (4.30) has been made; hence it is automatically fulfilled. We are therefore left with condition (4.67), where we note that
$$B_1\left([r]^{p1}\right)=2(\delta _{r,1}+\delta _{r,p})1,B_2\left([r]^{p1}\right)=2\delta _{r,1}1.$$
(4.68)
We conclude that finding all mixed multilinear relations of type (4.63) amounts to determining all sets of $`b_t^\nu `$ coefficients satisfying Eqs. (4.66)โ(4.68).
Once this has been done, it still remains to eliminate some dependent relations by taking into account identities such as
$$I_{r+1}Q_r=Q_1,r=2,3,\mathrm{},p,$$
(4.69)
$$I_tQ_1=Q_{t1},t=3,4,\mathrm{},p+1,$$
(4.70)
$$Q_{r_1}Q_{r_2}\mathrm{}Q_{r_p}Q_s^{}=I_tQ_{r_1}Q_{r_2}\mathrm{}Q_{r_p}Q_s^{},t=1,2,\mathrm{},p,$$
(4.71)
$`Q_{r_k}Q_{r_{k+1}}\mathrm{}Q_{r_p}Q_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_{k1}}=I_tQ_{r_k}Q_{r_{k+1}}\mathrm{}Q_{r_p}Q_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_{k1}},`$
$`k=2,3,\mathrm{},p,t=1,2,\mathrm{},p1,`$ (4.72)
$$Q_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_p}=I_tQ_s^{}Q_{r_1}Q_{r_2}\mathrm{}Q_{r_p},t=1,2,\mathrm{},p1,p+1.$$
(4.73)
By proceeding in this way for $`p=2`$, one gets the six relations given in Eqs. (4.51)โ(4.53). The $`p=3`$ case can be dealt with in a similar way, giving back the results of Khare (1993).
As a final point, let us note that there exists an alternative approach to PSSQM of order $`p`$, due to Beckers and Debergh (1990), wherein Eq. (4.3) is replaced by the cubic equation
$$[Q,[Q^{},Q]]=2Q,$$
(4.74)
while Eqs. (4.1) and (4.2) remain the same. We proved elsewhere (Quesne and Vansteenkiste, 1998) that in the $`p=2`$ case, Beckers-Debergh PSSQM algebra can only be realized by those $`๐^{(3)}(G(N))`$ algebras that simultaneously bosonize Rubakov-Spiridonov-Khare PSSQM algebra. For such a reason, we do not consider here that alternative approach to PSSQM of order $`p`$.
## 5 APPLICATION OF $`๐ช_\mathrm{๐}`$-EXTENDED OSCILLATOR ALGEBRAS TO PSEUDOSSQM
PseudoSSQM was introduced by Beckers et al. (1995a, b) (see also Beckers and Debergh (1995a, b)) in a study of relativistic vector mesons interacting with an external constant magnetic field, wherein the reality of energy eigenvalues was required. In the nonrelativistic limit, their theory leads to a pseudosupersymmetric oscillator Hamiltonian, which can be realized in terms of bosons and pseudofermions, where the latter are intermediate between standard fermions and parafermions of order two. It is then possible to formulate a pseudoSSQM, characterized by a pseudosupersymmetric Hamiltonian $``$ and pseudosupercharge operators $`Q`$, $`Q^{}`$, satisfying the relations
$$Q^2=0,$$
(5.1)
$$[,Q]=0,$$
(5.2)
$$QQ^{}Q=4c^2Q,$$
(5.3)
and their Hermitian conjugates, where $`c`$ is some real constant. The first two relations in Eqs. (5.1), (5.2) are the same as those occurring in SSQM, whereas the third one in Eq. (5.3) is similar to the multilinear relation valid in PSSQM of order two. Actually, for $`c=1`$ or 1/2, it is compatible with Eq. (4.3) or (4.74), respectively.
We will now show that the pseudoSSQM algebra (5.1)โ(5.3) can be realized in terms of the generators of $`๐^{(3)}(G(N))`$ in their bosonic Fock-space representation. For such a purpose, as in the $`p=2`$ PSSQM case (Quesne and Vansteenkiste, 1998), we shall start by assuming
$$Q=\underset{\nu =0}{\overset{2}{}}\left(\xi _\nu a+\eta _\nu a^{}\right)P_\nu ,$$
(5.4)
$$=H_0+\frac{1}{2}\underset{\nu =0}{\overset{2}{}}r_\nu P_\nu ,$$
(5.5)
where $`H_0`$ is the bosonic oscillator Hamiltonian (3.37) associated with $`๐^{(3)}(G(N))`$, $`\xi _\nu `$, $`\eta _\nu `$ are some complex constants, and $`r_\nu `$ some real ones, to be selected in such a way that Eqs. (5.1)โ(5.3) are satisfied.
Inserting the expression of $`Q`$, given in Eq. (5.4), into the first condition (5.1), we obtain some restrictions on the parameters $`\xi _\nu `$, $`\eta _\nu `$, leading to two sets of three independent solutions for $`Q`$. The solutions belonging to the first set are given by
$$Q=\left(\xi _{\mu +2}a+\eta _{\mu +2}a^{}\right)P_{\mu +2},$$
(5.6)
where $`\mu `$ takes some fixed, arbitrary value in the set $`\{0,1,2\}`$. Those belonging to the second set can be written as
$$Q^{}=\xi _{\mu +2}aP_{\mu +2}+\eta _\mu a^{}P_\mu ,$$
(5.7)
and can be obtained from the former by interchanging the roles of $`Q`$ and $`Q^{}`$ (and changing the $`\mu `$ value). They will be omitted here, since $`Q`$ and $`Q^{}`$ play a symmetrical role in the pseudoSSQM algebra (5.1)โ(5.3).
Considering next the second and third conditions in Eqs. (5.2) and (5.3), with Q given by Eq. (5.6) for some $`\mu `$ value, and the corresponding $``$ given by Eq. (5.5), we get the restrictions
$$\xi _{\mu +2}(2+\alpha _\mu +r_{\mu +1}r_{\mu +2})=0,$$
(5.8)
$$\eta _{\mu +2}(2\alpha _{\mu +1}+r_\mu r_{\mu +2})=0,$$
(5.9)
and
$$\left(|\xi _{\mu +2}|^2+|\eta _{\mu +2}|^2\right)\xi _{\mu +2}=4c^2\xi _{\mu +2},$$
(5.10)
$$\left(|\xi _{\mu +2}|^2+|\eta _{\mu +2}|^2\right)\eta _{\mu +2}=4c^2\eta _{\mu +2},$$
(5.11)
$`\xi _{\mu +2}\left[\left(|\xi _{\mu +2}|^2+|\eta _{\mu +2}|^2\right)(1+\alpha _{\mu +1})+|\eta _{\mu +2}|^2(1+\alpha _{\mu +2})\right]`$
$`=2c^2\xi _{\mu +2}(3+2\alpha _{\mu +1}+\alpha _{\mu +2}+r_{\mu +2}),`$ (5.12)
$$\eta _{\mu +2}|\eta _{\mu +2}|^2(1+\alpha _{\mu +2})=2c^2\eta _{\mu +2}(1+\alpha _{\mu +2}+r_{\mu +2}),$$
(5.13)
respectively.
Equations (5.8) and (5.9) have three independent solutions:
$$\xi _{\mu +2}0,\eta _{\mu +2}0,r_{\mu +1}=2\alpha _\mu +r_{\mu +2},r_\mu =2+\alpha _{\mu +1}+r_{\mu +2},$$
(5.14)
$$\xi _{\mu +2}0,\eta _{\mu +2}=0,r_{\mu +1}=2\alpha _\mu +r_{\mu +2},$$
(5.15)
$$\xi _{\mu +2}=0,\eta _{\mu +2}0,r_\mu =2+\alpha _{\mu +1}+r_{\mu +2}.$$
(5.16)
Since the third solution can be obtained from the second one by substituting $`Q^{}`$ for $`Q`$, and changing the $`\mu `$ value, we are only left with the first two solutions (5.14) and (5.15).
Introducing Eq. (5.14) into Eqs. (5.10)โ(5.13), we get the additional conditions
$$|\xi _{\mu +2}|=\sqrt{4c^2|\eta _{\mu +2}|^2},r_{\mu +2}=\frac{1}{2c^2}(1+\alpha _{\mu +2})\left(|\eta _{\mu +2}|^22c^2\right),$$
(5.17)
which define with Eq. (5.14) the first set of solutions of the pseudoSSQM algebra (5.1)โ(5.3). As we can fix the overall, arbitrary phase of $`Q`$ in such a way that $`\eta _{\mu +2}`$ is real and positive, we obtain for each $`\mu `$ value a two-parameter family of operators
$$Q(\eta _{\mu +2},\phi )=\left(\eta _{\mu +2}a^{}+e^{\mathrm{i}\phi }\sqrt{4c^2\eta _{\mu +2}^2}a\right)P_{\mu +2},$$
(5.18)
$$(\eta _{\mu +2})=N+\frac{1}{2}(2\gamma _{\mu +2}+r_{\mu +2}1)I+2P_{\mu +1}+P_{\mu +2},$$
(5.19)
where $`0<\eta _{\mu +2}<2|c|`$, $`0\phi <2\pi `$, and $`r_{\mu +2}`$ is given by Eq. (5.17). If we choose for instance $`\eta _{\mu +2}=\sqrt{2}|c|`$, and $`\phi =0`$, we get $`r_{\mu +2}=0`$, and
$$Q=c\sqrt{2}\left(a^{}+a\right)P_{\mu +2},$$
(5.20)
$$=N+\frac{1}{2}(2\gamma _{\mu +2}1)I+2P_{\mu +1}+P_{\mu +2}.$$
(5.21)
Note that this choice does not change $``$ in any significant way since it only produces an overall shift of its spectrum.
Introducing now Eq. (5.15) into Eqs. (5.10)โ(5.13) we get instead the additional conditions
$$|\xi _{\mu +2}|=2|c|,r_{\mu +2}=1\alpha _{\mu +2},$$
(5.22)
which define with Eq. (5.15) a second set of solutions of the pseudoSSQM algebra (5.1)โ(5.3). Choosing this time the overall, arbitrary phase of $`Q`$ in such a way that $`\xi _{\mu +2}`$ is real and positive, we obtain for each $`\mu `$ value a one-parameter family of operators
$$Q=2|c|aP_{\mu +2},$$
(5.23)
$$(r_\mu )=N+\frac{1}{2}(2\gamma _{\mu +2}\alpha _{\mu +2})I+\frac{1}{2}(1\alpha _{\mu +1}+\alpha _{\mu +2}+r_\mu )P_\mu +P_{\mu +1},$$
(5.24)
where the parameter $`r_\mu `$ does change the Hamiltonian spectrum in a significant way.
The pseudosupersymmetric Hamiltonian, corresponding to the first solution (5.20), (5.21), coincides with the $`p=2`$ parasupersymmetric Hamiltonian previously obtained (Quesne and Vansteenkiste, 1998), and defined for arbitrary $`p`$ in Eq. (4.34) of the present work (but the respective charges are of course different). Its spectrum and its ground-state energy are therefore given by Eqs. (4.35), (4.36), and by Eq. (4.37), respectively.
On the contrary, the pseudosupersymmetric Hamiltonian $`(r_\mu )`$, corresponding to the second solution (5.23), (5.24), is new, and its spectrum is given by
$$_{3k+\nu }=3k+\frac{1}{2}(2\gamma _{\mu +2}\alpha _{\mu +2}+2\mu 2),\mathrm{if}\nu =0,1,\mathrm{},\mu 1,$$
(5.25)
$$_{3k+\mu }=3k+\frac{1}{2}(2\gamma _\mu +r_\mu +2\mu +1),$$
(5.26)
$$_{3k+\nu }=3k+\frac{1}{2}(2\gamma _{\mu +2}\alpha _{\mu +2}+2\mu +4),\mathrm{if}\nu =\mu +1,\mu +2,\mathrm{},2.$$
(5.27)
Its levels are therefore equally spaced only if $`r_\mu =(\alpha _{\mu +1}\alpha _{\mu +2}+3)\mathrm{mod}\mathrm{\hspace{0.17em}6}`$. If $`r_\mu `$ is small enough, the ground state is nondegenerate, and its energy is negative for $`\mu =1`$, or may have any sign for $`\mu =0`$ or 2. On the contrary, if $`r_\mu `$ is large enough, the ground state remains nondegenerate with a vanishing energy in the former case, while it becomes twofold degenerate with a positive energy in the latter. For some intermediate $`r_\mu `$ value, one gets a two or threefold degenerate ground state with a vanishing or positive energy, respectively.
## 6 APPLICATION OF $`๐ช_\mathrm{๐}`$-EXTENDED OSCILLATOR ALGEBRAS TO OSSQM OF ORDER TWO
OSSQM of arbitrary order $`p`$ was developed by Khare et al. (1993a), by combining standard bosons with orthofermions of order $`p`$. The latter had been previously introduced by Mishra and Rajasekaran (1991a, b), by replacing Pauliโs exclusion principle by a new, more stringent one. OSSQM is formulated in terms of an orthosupersymmetric Hamiltonian $``$, and $`p`$ orthosupercharge operators $`Q_r`$, $`Q_r^{}`$, $`r=1`$, 2, โฆ, $`p`$, satisfying the relations
$$Q_rQ_s=0,$$
(6.1)
$$[,Q_r]=0,$$
(6.2)
$$Q_rQ_s^{}+\delta _{r,s}\underset{t=1}{\overset{p}{}}Q_t^{}Q_t=2\delta _{r,s},$$
(6.3)
and their Hermitian conjugates, where $`r`$ and $`s`$ run over 1, 2, โฆ, $`p`$.
We plan to show that for $`p=2`$, the OSSQM algebra (6.1)โ(6.3) can be realized in terms of the generators of $`๐^{(3)}(G(N))`$ in their bosonic Fock-space representation. For such a purpose, let us set
$$Q_1=\underset{\nu =0}{\overset{2}{}}\left(\xi _\nu a+\eta _\nu a^{}\right)P_\nu ,$$
(6.4)
$$Q_2=\underset{\nu =0}{\overset{2}{}}\left(\zeta _\nu a+\rho _\nu a^{}\right)P_\nu ,$$
(6.5)
$$=H_0+\frac{1}{2}\underset{\nu =0}{\overset{2}{}}r_\nu P_\nu ,$$
(6.6)
where we now have at our disposal four types of complex constants $`\xi _\nu `$, $`\eta _\nu `$, $`\zeta _\nu `$, $`\rho _\nu `$, and one of real ones $`r_\nu `$, to adjust in order that Eqs. (6.1)โ(6.3) be satisfied.
Let us first consider Eq. (6.1) for $`r=s=1`$, 2. From the study carried out in Section 5, we know that for each $`r`$ in the set $`\{1,2\}`$, the equation $`Q_r^2=0`$ admits two different types of solutions, given in Eqs. (5.6) and (5.7), respectively, and connected by the symmetry $`QQ^{}`$. In the present case, we have to distinguish them, since the OSSQM algebra (6.1)โ(6.3) is not invariant under such a symmetry. Hence, for the couple of orthosupersymmetric charges $`(Q_1,Q_2)`$, we get seven types of solutions of $`Q_1^2=Q_2^2=0`$, namely $`Q_1`$ and $`Q_2`$ may be both of type $`Q`$, or $`Q^{}`$, with the same or adjacent $`\mu `$ values, or $`Q_1`$ is of type $`Q`$ corresponding to a given $`\mu `$ value, and $`Q_2`$ of type $`Q^{}`$ corresponding to $`\mu `$, $`\mu +1`$, or $`\mu +2`$. Here, we take into account the fact that the algebra (6.1)โ(6.3) is invariant under the exchange $`Q_1Q_2`$.
Imposing next Eqs. (6.1) and (6.3) for $`rs`$, i.e., $`Q_1Q_2=Q_2Q_1=Q_1Q_2^{}=0`$, we obtain that those seven cases for $`(Q_1,Q_2)`$ actually reduce to two, given by
$$Q_1=\xi _{\mu +2}aP_{\mu +2}+\eta _\mu a^{}P_\mu ,Q_2=\zeta _{\mu +2}aP_{\mu +2}+\rho _\mu a^{}P_\mu ,$$
(6.7)
and
$$Q_1=\xi _{\mu +2}aP_{\mu +2},Q_2=\rho _\mu a^{}P_\mu ,$$
(6.8)
respectively, where for the first one, we have the additional conditions
$$\xi _{\mu +2}\zeta _{\mu +2}^{}+\eta _\mu \rho _\mu ^{}=0(\xi _{\mu +2},\eta _\mu 0),$$
(6.9)
$$\alpha _{\mu +1}=1.$$
(6.10)
Note that the latter is compatible with conditions (3.34) for the existence of the bosonic Fock-space representation only for $`\mu =0`$ and $`\mu =1`$.
Equation (6.2) now leads to the same conditions for both choices (6.7) and (6.8), namely
$$r_\mu =4+\alpha _{\mu +1}+r_{\mu +2},r_{\mu +1}=2\alpha _\mu +r_{\mu +2}.$$
(6.11)
It only remains to impose Eq. (6.3) for $`r=s=1`$, 2. For the first couple of operators $`(Q_1,Q_2)`$, given in Eqs. (6.7), (6.9), and (6.10), we obtain the additional restrictions
$$|\xi _{\mu +2}|^2+|\eta _\mu |^2=2,|\zeta _{\mu +2}|^2=|\eta _\mu |^2,|\rho _\mu |^2=|\xi _{\mu +2}|^2,\xi _{\mu +2}\eta _\mu ^{}+\zeta _{\mu +2}\rho _\mu ^{}=0,$$
(6.12)
and
$$r_\mu =1+\alpha _\mu ,r_{\mu +1}=0,r_{\mu +2}=1\alpha _{\mu +2}.$$
(6.13)
Combining Eqs. (6.9) and (6.12), we get
$$\xi _{\mu +2}=|\xi _{\mu +2}|e^{\mathrm{i}\alpha },\eta _\mu =\sqrt{2|\xi _{\mu +2}|^2}e^{\mathrm{i}\beta },$$
(6.14)
$$\zeta _{\mu +2}=\sqrt{2|\xi _{\mu +2}|^2}e^{\mathrm{i}(\alpha \beta +\gamma )},\rho _\mu =|\xi _{\mu +2}|e^{\mathrm{i}\gamma },$$
(6.15)
where $`0<|\xi _{\mu +2}|<\sqrt{2}`$, and $`0\alpha ,\beta ,\gamma <2\pi `$. In addition, we find that Eqs. (6.10), (6.11), and (6.13) are compatible, and can be combined into the relations
$$r_\mu =1+\alpha _\mu ,r_{\mu +1}=0,r_{\mu +2}=2+\alpha _\mu ,\alpha _{\mu +1}=1.$$
(6.16)
Choosing the overall, arbitrary phases of $`Q_1`$ and $`Q_2`$ in such a way that $`\xi _{\mu +2}`$ and $`\rho _\mu `$ are real and positive, and setting $`\beta =\phi `$, we obtain, for $`\mu =0`$ or 1, a two-parameter family of solutions of Eqs. (6.1)โ(6.3),
$$Q_1(\xi _{\mu +2},\phi )=\xi _{\mu +2}aP_{\mu +2}+e^{\mathrm{i}\phi }\sqrt{2\xi _{\mu +2}^2}a^{}P_\mu ,$$
(6.17)
$$Q_2(\xi _{\mu +2},\phi )=e^{\mathrm{i}\phi }\sqrt{2\xi _{\mu +2}^2}aP_{\mu +2}+\xi _{\mu +2}a^{}P_\mu ,$$
(6.18)
$$=N+\frac{1}{2}(2\gamma _{\mu +1}1)I+2P_\mu +P_{\mu +1},$$
(6.19)
where $`0<\xi _{\mu +2}<\sqrt{2}`$, $`0\phi <2\pi `$, and $`\alpha _{\mu +1}=1`$.
For the second couple of operators $`(Q_1,Q_2)`$, given in Eq. (6.8), Eq. (6.3) with $`r=s=1`$, 2 leads to the conditions
$$|\xi _{\mu +2}|^2=|\rho _\mu |^2=2,$$
(6.20)
and to Eqs. (6.10) and (6.13). Hence, with an appropriate choice of phases, we obtain Eqs. (6.17)โ(6.19) with $`\xi _{\mu +2}=\sqrt{2}`$. We conclude that the most general solution of the OSSQM algebra (6.1)โ(6.3) that can be written in the form (6.4)โ(6.6) is given by Eqs. (6.17)โ(6.19), where $`\mu \{0,1\}`$, $`0<\xi _{\mu +2}\sqrt{2}`$, $`0\phi <2\pi `$, and $`\alpha _{\mu +1}=1`$.
The orthosupersymmetric Hamiltonian $``$ in Eq. (6.19) is independent of the parameters $`\xi _{\mu +2}`$, $`\phi `$. All the levels of its spectrum are equally spaced. For $`\mu =0`$, they are threefold degenerate, since
$$_{3k}=_{3k+1}=_{3k+2}=3k+\frac{1}{2}(2\gamma _1+3).$$
(6.21)
OSSQM is therefore broken, and the ground-state energy
$$_0=_1=_2=\frac{1}{2}(2\gamma _1+3)=\alpha _0+1$$
(6.22)
is positive. On the contrary, for $`\mu =1`$, only the excited states are threefold degenerate, since
$$_{3(k+1)}=_{3k+1}=_{3k+2}=3k+\frac{1}{2}(2\gamma _2+5).$$
(6.23)
OSSQM is then unbroken, and the ground-state energy
$$_0=\frac{1}{2}(2\gamma _21)=\frac{1}{2}(\alpha _2+1)$$
(6.24)
vanishes. Such results agree with the general conclusions of Khare et al. (1993a).
For $`p`$ values greater than two, the OSSQM algebra (6.1)โ(6.3) becomes rather complicated because the number of equations to be fulfilled increases considerably. A glance at the 18 independent conditions for $`p=3`$ led us to the conclusion that the $`๐^{(4)}(G(N))`$ algebra is not rich enough to contain operators satisfying Eqs. (6.1)โ(6.3). Contrary to what happens for PSSQM, for OSSQM the $`p=2`$ case is therefore not representative of the general one.
## 7 SOME DEFORMED $`๐ช_๐`$-EXTENDED OSCILLATOR ALGEBRAS
The purpose of the present section is to construct some deformations of the $`C_\lambda `$-extended oscillator algebras $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, subject to the condition that they admit three Casimir operators analogous to $`๐_1`$, $`๐_2`$, $`๐_3`$, defined in Eqs. (2.10)โ(2.12).
Let us consider a class of algebras generated by $`I`$, $`a^{}`$, $`a=\left(a^{}\right)^{}`$, $`N=N^{}`$, $`P_\mu ^{}=P_\mu ^{}`$, $`\mu =0`$, 1, $`\mathrm{}`$$`\lambda 1`$, satisfying the defining relations (2.5)โ(2.7) of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, except for the commutator of $`a`$ and $`a^{}`$ in Eq. (2.7), which is replaced by the quommutator (or $`q`$-deformed commutator)
$$[a,a^{}]_qaa^{}qa^{}a=H(N)+K(N)\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,$$
(7.1)
where $`q\text{}^+`$, $`\alpha _\mu \text{}`$, and $`H(N)`$, $`K(N)`$ are some real, analytic functions of $`N`$.
The operators $`๐_1`$, $`๐_2`$ of Eqs. (2.10), (2.11) remain invariants of the new algebras. We will determine the constraints that the existence of a third Casimir operator of the type
$$\stackrel{~}{๐}_3=q^N\left(D(N)+E(N)\underset{\mu =0}{\overset{\lambda 1}{}}\beta _\mu P_\mu a^{}a\right)$$
(7.2)
imposes on $`H(N)`$ and $`K(N)`$, assuming that Eq. (2.9) is the only relation satisfied by the $`\alpha _\mu `$โs. Here $`\beta _\mu `$, $`\mu =0`$, 1, โฆ, $`\lambda 1`$, and $`D(N)`$, $`E(N)`$ are assumed to be some real constants, and some real, analytic functions of $`N`$, respectively. In the case of the undeformed algebras $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, one has $`q=1`$, $`H(N)=K(N)=I`$, and $`\stackrel{~}{๐}_3`$ reduces to $`๐_3`$, given in Eq. (2.12), with $`D(N)=N`$, $`E(N)=I`$, and $`\beta _\mu `$ defined by Eq. (2.13) in terms of the $`\alpha _\mu `$โs.
In the realization (2.16), the deformed algebras, defined by Eqs. (2.5), (2.6), (2.9), and (7.1), reduce to GDOAs $`๐_q^{(\lambda )}(G(N))`$, with $`q1`$ and $`G(N)`$ given by the right-hand side of Eq. (7.1). Then $`\stackrel{~}{๐}_3`$ reduces to the standard Casimir operator $`\stackrel{~}{๐}`$ of such algebras, and $`F(N)=D(N)+E(N)_{\mu =0}^{\lambda 1}\beta _\mu P_\mu `$ becomes the GDOA structure function, satisfying the equation $`F(N+1)qF(N)=G(N)`$ (Katriel and Quesne, 1996; Quesne and Vansteenkiste, 1996, 1997).
Going back to the general case, we note that since $`\stackrel{~}{๐}_3`$ is a Hermitian operator commuting with $`N`$ and $`P_\mu `$, we only have to impose the condition $`[\stackrel{~}{๐}_3,a]=0`$. By using the defining relations, it is easy to show that the latter is equivalent to the two functional equations
$$D(N+1)qD(N)=H(N),$$
(7.3)
$$E(N+1)\beta _{\mu +1}qE(N)\beta _\mu =K(N)\alpha _\mu ,\mu =0,1,\mathrm{},\lambda 1,$$
(7.4)
where we assume as usual $`\beta _\lambda =\beta _0`$. Equation (7.3) is similar to the equation appearing in the construction of $`\stackrel{~}{๐}`$ for GDOAs with $`q1`$ (Katriel and Quesne, 1996; Quesne and Vansteenkiste, 1996, 1997), while Eq. (7.4) is a new functional equation, whose solutions will now be determined.
For such a purpose, let us consider the following nonhomogeneous system of $`\lambda `$ linear equations in $`\lambda `$ unknowns $`\beta _\mu `$, $`\mu =0`$, 1, โฆ, $`\lambda 1`$,
$$qE(x)\beta _\mu +E(x+1)\beta _{\mu +1}=K(x)\alpha _\mu ,\mu =0,1,\mathrm{},\lambda 1,$$
(7.5)
$$\beta _\lambda \beta _0,$$
(7.6)
where $`x`$ is some real variable.
If the determinant of its coefficient matrix is nonvanishing, i.e., if
$$[E(x+1)]^\lambda [qE(x)]^\lambda 0,$$
(7.7)
or, equivalently,
$$E(x)bq^x,$$
(7.8)
and
$$E(x)b^{}(q)^x,\text{if }\lambda \text{ is even},$$
(7.9)
where $`b`$, $`b^{}`$ are some real, nonvanishing constants, then the system has one and only one solution, given by
$$\beta _\mu =\frac{[qE(x)]^{\lambda 1}K(x)}{[E(x+1)]^\lambda [qE(x)]^\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}\left(\frac{E(x+1)}{qE(x)}\right)^\nu \alpha _{\mu +\nu },\mu =0,1,\mathrm{},\lambda 1.$$
(7.10)
Since, by definition, $`\beta _\mu `$, $`\mu =0`$, 1, โฆ, $`\lambda 1`$, are constants, the functions $`E(x)`$ and $`K(x)`$ should be chosen in such a way that the dependence on $`x`$ disappears on the right-hand side of Eq. (7.10).
Let us first consider $`\beta _0`$. By using Eq. (2.9) to express $`\alpha _0`$ in terms of $`\alpha _1`$, $`\alpha _2`$, โฆ, $`\alpha _{\lambda 1}`$, $`\beta _0`$ can be rewritten as
$$\beta _0=\frac{[qE(x)]^{\lambda 1}K(x)}{[E(x+1)]^\lambda [qE(x)]^\lambda }\underset{\nu =1}{\overset{\lambda 1}{}}\left[\left(\frac{E(x+1)}{qE(x)}\right)^\nu 1\right]\alpha _\nu .$$
(7.11)
Since $`\alpha _1`$, $`\alpha _2`$, โฆ, $`\alpha _{\lambda 1}`$ are assumed to be independent, the coefficient of each of them on the right-hand side of Eq. (7.11) should reduce to some real constant, which we denote by $`e_\nu `$, $`\nu =1`$, 2, โฆ, $`\lambda 1`$. Hence we get the system of equations
$$\frac{1}{e_1}\left(\frac{E(x+1)}{qE(x)}1\right)=\frac{[E(x+1)]^\lambda [qE(x)]^\lambda }{[qE(x)]^{\lambda 1}K(x)},$$
(7.12)
$$\frac{1}{e_1}\left(\frac{E(x+1)}{qE(x)}1\right)=\frac{1}{e_\nu }\left[\left(\frac{E(x+1)}{qE(x)}\right)^\nu 1\right],\nu =2,3,\mathrm{},\lambda 1,$$
(7.13)
to determine the constraints on $`E(x)`$ and $`K(x)`$.
For $`\lambda =2`$, we are only left with the first equation (7.12), yielding the constraint
$$K(x)=e_1[E(x+1)+qE(x)].$$
(7.14)
Introducing the latter into Eq. (7.10), and using Eq. (2.9) again, we obtain
$$\beta _\mu =e_1\alpha _\mu ,\mu =0,1,$$
(7.15)
which are constants as it should be. Incorporating the constant $`e_1`$ into the $`E(x)`$ definition, we conclude that the algebras defined by Eqs. (2.5), (2.6) with $`\lambda =2`$, and
$$[a,a^{}]_q=H(N)+[E(N+1)+qE(N)](\alpha _0P_0+\alpha _1P_1),$$
(7.16)
where $`\alpha _0`$, $`\alpha _1`$ satisfy Eq. (2.9), $`H(N)`$ is arbitrary, and $`E(N)(\pm q)^N`$, admit the three Casimir operators (2.10), (2.11), and
$$\stackrel{~}{๐}_3=q^N\left[D(N)E(N)(\alpha _0P_0+\alpha _1P_1)a^{}a\right],$$
(7.17)
where $`D(N)`$ is some solution of Eq. (7.3). By choosing that solution for which $`D(0)=\alpha _0E(0)`$, $`\stackrel{~}{๐}_3`$ vanishes in the bosonic Fock-space representation.
For $`\lambda >2`$, Eq. (7.13) for $`\nu =2`$ yields the constraint
$$E(x+1)=\left(\frac{e_2}{e_1}1\right)qE(x),$$
(7.18)
whose solution is given by
$$E(x)=bk^x,$$
(7.19)
where $`b`$ is some real constant, and $`k\left(e_1^1e_21\right)q`$. From Eqs. (7.8) and (7.9), it follows that for any $`\lambda `$, $`kq`$, and in addition for even $`\lambda `$, $`kq`$. Equation (7.12) then provides the expression of $`K(x)`$,
$$K(x)=Bk^x,$$
(7.20)
where $`Be_1bq^{2\lambda }\left(k^\lambda q^\lambda \right)/(kq)`$, while for the remaining $`\nu `$ values, Eq. (7.13) leads to the conditions
$$e_\nu =e_1q^{1\nu }\frac{k^\nu q^\nu }{kq},\nu =2,3,\mathrm{},\lambda 1.$$
(7.21)
Hence, from Eq. (7.10), $`\beta _\mu `$ is given by
$$\beta _\mu =\frac{Bq^{\lambda 1}}{b(k^\lambda q^\lambda )}\underset{\nu =0}{\overset{\lambda 1}{}}\left(\frac{k}{q}\right)^\nu \alpha _{\mu +\nu },\mu =0,1,\mathrm{},\lambda 1,$$
(7.22)
and therefore reduces to some constant as it should be. We conclude that for $`\lambda >2`$, the algebras defined by Eqs. (2.5), (2.6), and
$$[a,a^{}]_q=H(N)+Bk^N\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,$$
(7.23)
where $`H(N)`$ and $`B`$ are arbitrary, $`\alpha _\mu `$ satisfies Eq. (2.9), $`kq`$ for any $`\lambda `$, and $`kq`$ for even $`\lambda `$, admit the three Casimir operators (2.10), (2.11), and
$$\stackrel{~}{๐}_3=q^N\left\{D(N)+\frac{Bq^{\lambda 1}}{k^\lambda q^\lambda }k^N\underset{\mu =0}{\overset{\lambda 1}{}}\left[\underset{\nu =0}{\overset{\lambda 1}{}}\left(\frac{k}{q}\right)^\nu \alpha _{\mu +\nu }\right]P_\mu a^{}a\right\},$$
(7.24)
where $`D(N)`$ is some solution of Eq. (7.3). By choosing that solution for which $`D(0)=Bq^{\lambda 1}\left(k^\lambda q^\lambda \right)^1_{\nu =0}^{\lambda 1}(k/q)^\nu \alpha _\nu `$, $`\stackrel{~}{๐}_3`$ vanishes in the bosonic Fock-space representation.
It remains to consider the cases where the coefficient matrix of system (7.5), (7.6) has a vanishing determinant. If $`E(x)=bq^x`$, where $`b`$ is some real constant, then Eqs. (7.5) and (7.6) become
$$\beta _\mu +\beta _{\mu +1}=(bq)^1\frac{K(x)}{q^x}\alpha _\mu ,\mu =0,1,\mathrm{},\lambda 1,$$
(7.25)
$$\beta _\lambda \beta _0.$$
(7.26)
Since the $`\beta _\mu `$โs are constants, we obtain
$$K(x)=Bq^x,$$
(7.27)
where $`B`$ is some real constant, and therefore
$$\beta _\mu =\frac{B}{bq}\underset{\nu =0}{\overset{\mu 1}{}}\alpha _\nu +\beta _0,\mu =1,2,\mathrm{},\lambda 1.$$
(7.28)
We conclude that the algebras defined by Eqs. (2.5), (2.6), and
$$[a,a^{}]_q=H(N)+Bq^N\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,$$
(7.29)
where $`H(N)`$ and $`B`$ are arbitrary, and $`\alpha _\mu `$ satisfies Eq. (2.9), admit the three Casimir operators (2.10), (2.11), and
$$\stackrel{~}{๐}_3=q^N\left[D(N)+Bq^{N1}\underset{\mu =1}{\overset{\lambda 1}{}}\left(\underset{\nu =0}{\overset{\mu 1}{}}\alpha _\nu \right)P_\mu a^{}a\right],$$
(7.30)
where we have set $`\beta _0=0`$ (thereby eliminating a multiple of the unit operator), and $`D(N)`$ is some solution of Eq. (7.3). By choosing that solution for which $`D(0)=0`$, $`\stackrel{~}{๐}_3`$ vanishes in the bosonic Fock-space representation.
Finally, if $`\lambda `$ is even, and $`E(x)=b(q)^x`$, where $`b`$ is some real constant, then Eqs. (7.5) and (7.6) become
$$\beta _\mu +\beta _{\mu +1}=(bq)^1\frac{K(x)}{(q)^x}\alpha _\mu ,\mu =0,1,\mathrm{},\lambda 1,$$
(7.31)
$$\beta _\lambda \beta _0.$$
(7.32)
The $`\beta _\mu `$ constancy implies again that
$$K(x)=B(q)^x,$$
(7.33)
where $`B`$ is some real constant. Equation (7.31) is then equivalent to
$$\beta _0+\beta _1=\frac{B}{bq}\alpha _0,$$
(7.34)
$$\beta _{\mu +2}\beta _\mu =\frac{B}{bq}(\alpha _{\mu +1}\alpha _\mu ),\mu =0,1,\mathrm{},\lambda 2.$$
(7.35)
The solution of Eqs. (7.34) and (7.35) is given by
$$\beta _\mu =\frac{B}{bq}\left(\underset{\nu =0}{\overset{(\mu 2)/2}{}}\alpha _{2\nu +1}\underset{\nu =0}{\overset{(\mu 2)/2}{}}\alpha _{2\nu }\right)+\beta _0,\text{if }\mu \text{ is even},$$
(7.36)
$$\beta _\mu =\frac{B}{bq}\left(\underset{\nu =0}{\overset{(\mu 1)/2}{}}\alpha _{2\nu }\underset{\nu =0}{\overset{(\mu 3)/2}{}}\alpha _{2\nu +1}\right)\beta _0,\text{if }\mu \text{ is odd}.$$
(7.37)
Condition (7.32) is consistent with Eq. (7.36) if and only if we impose that $`_{\nu =0}^{(\lambda 2)/2}\alpha _{2\nu +1}=_{\nu =0}^{(\lambda 2)/2}\alpha _{2\nu }`$, or by taking Eq. (2.9) into account, $`_{\nu =0}^{(\lambda 2)/2}\alpha _{2\nu }=0`$. Since we have assumed that the $`\alpha _\mu `$โs do not satisfy any extra relation apart from Eq. (2.9), the case $`E(x)=b(q)^x`$ has to be rejected.
We therefore found altogether three deformed $`C_\lambda `$-extended oscillator algebras admitting three Casimir operators $`๐_1`$, $`๐_2`$, $`\stackrel{~}{๐}_3`$. They correspond to Eqs. (7.16) and (7.17), (7.23) and (7.24), (7.29) and (7.30), respectively.
The deformed Calogero-Vasiliev algebra introduced by Brzeziลski et al. (1993), for which
$$[a,a^{}]_q=q^N(1+2\alpha K),K=(1)^N,$$
(7.38)
is a special case of Eq. (7.16), corresponding to
$$H(N)=q^N,E(N)=\frac{2q^N}{q+q^1},\alpha _0=\alpha _1=\alpha .$$
(7.39)
From Eq. (7.3), we obtain
$$D(N)=\frac{q^Nq^N}{qq^1}+\frac{2\alpha q^N}{q+q^1},$$
(7.40)
so that the Casimir operator (7.17) becomes
$$\stackrel{~}{๐}_3=q^N\left(\frac{q^Nq^N}{qq^1}+\frac{2\alpha (q^Nq^NK)}{q+q^1}a^{}a\right).$$
(7.41)
In a given unirrep, whose basis states are given by Eq. (3.1), and satisfy relations similar to Eqs. (3.2) and (3.3) with $`๐=๐_3`$ replaced by $`\stackrel{~}{๐}_3`$, we obtain from Eq. (7.41) that $`\lambda _n`$ can be expressed as
$$\lambda _n=q^{n_0+n}c+\frac{q^{n_0+n}q^{n_0n}}{qq^1}+2\alpha \frac{q^{n_0+n}(q)^{n_0n}}{q+q^1},$$
(7.42)
or
$$\lambda _n=q^n\lambda _0+q^{n_0}\left(\frac{q^nq^n}{qq^1}+B\frac{q^n(q)^n}{q+q^1}\right),B2\alpha (1)^{n_0}.$$
(7.43)
This equation is consistent with Eq. (14) of Kosiลski et al. (1997), wherein the representations of the deformed Calogero-Vasiliev algebra were studied. Note that this result holds although there are some slight discrepancies in the algebra definition between Kosiลski et al. (1997) and the present work, and the Casimir operator $`\stackrel{~}{๐}_3`$ was not considered in the former.
Some interesting special cases of the algebras corresponding to Eqs. (7.23) and (7.29) are obtained for $`q=1`$, $`k1`$, $`k1`$ (if $`\lambda `$ is even), and $`k=1`$, $`q1`$, $`q1`$ (if $`\lambda `$ is even) for the former, and $`q=1`$ for the latter.
## 8 CONCLUSION
In the present paper, we studied some mathematical properties of $`C_\lambda `$-extended oscillator algebras $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$. We constructed Casimir operators, and used them to provide a complete unirrep classification under the assumption that the number operator spectrum is nondegenerate. We established that only BFB and FD unirreps occur, and showed that the unirreps of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$ can be related to those of its GDOA realization $`๐^{(\lambda )}(G(N))`$.
In addition, we looked for some deformations of $`๐_{\alpha _0\alpha _1\mathrm{}\alpha _{\lambda 2}}^{(\lambda )}`$, subject to the condition that they admit Casimir operators analogous to those of the undeformed algebras. We found three new types of algebras, defined in Eqs. (7.16) and (7.17), (7.23) and (7.24), (7.29) and (7.30), respectively. The first one includes the Brzeziลski et al. (1993) deformation of the Calogero-Vasiliev algebra (Vasiliev, 1991; Polychronakos, 1992; Brink et al., 1992; Brink and Vasiliev, 1993) as a special case.
Furthermore, we established that the bosonic Fock-space realization of $`๐^{(\lambda )}(G(N))`$ yields a convenient bosonization of several SSQM variants: PSSQM of order $`p=\lambda 1`$ for any $`\lambda `$, as well as pseudoSSQM, and OSSQM of order two for $`\lambda =3`$. In the former case, we provided a full analysis of the problem, including the construction of the $`p`$ independent conserved parasupercharges, and $`p`$ bosonic constants admitted by the parasupersymmetric Hamiltonian. Such results generalize those already known for standard SSQM (Brzeziลski et al., 1993; Plyushchay, 1996a, b). In the OSSQM case, however, it was not possible to extend the results to $`p`$ values greater than two in the $`C_\lambda `$-extended oscillator algebra context.
There remain some interesting open questions for future study. Apart from those mentioned in Section 1, we would like to mention here two of them. The first one is to further study deformations both from theoretical and applied viewpoints. Generalizing, for instance, the Macfarlane (1994) deformation of the Calogero-Vasiliev algebra would be an interesting topic. The second issue is to construct some GDOA, whose structure would be rich enough to enable the OSSQM bosonization to be carried out for $`p>2`$.
## ACKNOWLEDGMENT
One of the authors (CQ) is a Research Director of the National Fund for Scientific Research (FNRS), Belgium. The other (NV) is a Scientific Associate of the Inter-University Institute for Nuclear Sciences (IISN), Belgium.
## FOOTNOTES
<sup>1</sup>In both the oscillator and Heisenberg algebras, the creation and annihilation operators $`a^{}`$, $`a`$ are considered as generators, but in the former the number operator $`N`$ appears as an additional independent generator, whereas in the latter it is defined in terms of $`a^{}`$, $`a`$ as $`Na^{}a`$.
<sup>2</sup>In a recent study (Guichardet, 1998), the assumption that the spectrum of $`N`$ is nondegenerate has been lifted for the Arik-Coon GDOA (Arik and Coon, 1976; Kuryshkin, 1980), but it has been shown that this condition is automatically fulfilled.
## REFERENCES
Arik, M., and Coon, D. D. (1976). Journal of Mathematical Physics, 17, 524.
Bagchi, B. (1994). Physics Letters A, 189, 439.
Bagchi, B., Biswas, S. N., Khare, A., and Roy, P. K. (1997). Pramana \- Journal of Physics, 49, 199.
Beckers, J., and Debergh, N. (1990). Nuclear Physics B, 340, 767.
Beckers, J., and Debergh, N. (1995a). International Journal of Modern Physics A, 10, 2783.
Beckers, J., and Debergh, N. (1995b). In Second International Workshop on Harmonic Oscillators, Cocoyoc, Morelos, Mexico, March 23โ25, 1994, D. Han and K. B. Wolf, eds., NASA Conference Publication 3286, NASA Goddard Space Flight Center, Greenbelt, Maryland, p. 313.
Beckers, J., Debergh, N., and Nikitin, A. G. (1995a). Fortschritte der Physik, 43, 67.
Beckers, J., Debergh, N., and Nikitin, A. G. (1995b). Fortschritte der Physik, 43, 81.
Beckers, J., Debergh, N., and Nikitin, A. G. (1997). International Journal of Theorical Physics, 36, 1991.
Biedenharn, L. C. (1989). Journal of Physics A, 22, L873.
Bogoliubov, N. M., Rybin, A. V., and Timonen, J. (1994). Journal of Physics A, 27, L363.
Bonatsos, D. (1992). Journal of Physics A, 25, L101.
Bonatsos, D., and Daskaloyannis, C. (1992a). Physics Letters B, 278, 1.
Bonatsos, D., and Daskaloyannis, C. (1992b). Physical Review A, 46, 75.
Bonatsos, D., and Daskaloyannis, C. (1993a). Physics Letters B, 307, 100.
Bonatsos, D., and Daskaloyannis, C. (1993b). Chemical Physics Letters, 203, 150.
Bonatsos, D., and Daskaloyannis, C. (1993c). Physical Review A, 48, 3611.
Bonatsos, D., Daskaloyannis, C., and Kokkotas, K. (1993). Physical Review A, 48, R3407.
Bonatsos, D., Daskaloyannis, C., and Kokkotas, K. (1994). Physical Review A, 50, 3700.
Brink, L., Hansson, T. H., and Vasiliev, M. A. (1992). Physics Letters B, 286, 109.
Brink, L., and Vasiliev, M. A. (1993). Modern Physics Letters A, 8, 3585.
Brzeziลski, T., Egusquiza, I. L., and Macfarlane, A. J. (1993). Physics Letters B, 311, 202.
Calogero, F. (1969a). Journal of Mathematical Physics, 10, 2191.
Calogero, F. (1969b). Journal of Mathematical Physics, 10, 2197.
Calogero, F. (1971). Journal of Mathematical Physics, 12, 419.
Calogero, F., and Marchioro, C. (1974). Journal of Mathematical Physics, 15, 1425.
Chang, Z., Guo, H. Y., and Yan, H. (1991). Physics Letters A, 156, 192.
Chang, Z., and Yan, H. (1991a). Physics Letters A, 158, 242.
Chang, Z., and Yan, H. (1991b). Physical Review A, 43, 6043.
Chang, Z., and Yan, H. (1991c). Physical Review A, 44, 7405.
Cornwell, J. F. (1984). Group Theory in Physics, Vol. 1, Academic Press, New York, p. 117.
Daskaloyannis, C. (1991). Journal of Physics A, 24, L789.
Daskaloyannis, C. (1992). Journal of Physics A, 25, 2261.
Eleonsky, V. M., and Korolev, V. G. (1995). Journal of Physics A, 28, 4973.
Eleonsky, V. M., Korolev, V. G., and Kulagin, N. E. (1994). Chaos, 4, 583.
Eleonsky, V. M., Korolev, V. G., and Kulagin, N. E. (1995). Doklady Rossiiskoi Akademii Nauk, 342, 1.
Fairlie, D. B., and Nuyts, J. (1994). Journal of Mathematical Physics, 35, 3794.
Fairlie, D. B., and Zachos, C. K. (1991). Physics Letters B, 256, 43.
Fivel, D. I. (1990). Physical Review Letters, 65, 3361.
Green, H. S. (1953). Physical Review, 90, 270.
Greenberg, O. W. (1990). Physical Review Letters, 64, 705.
Greenberg, O. W. (1991). Physical Review D, 43, 4111.
Guichardet, A. (1998). Journal of Mathematical Physics, 39, 4965.
Hayashi, T. (1990). Communications in Mathematical Physics, 127, 129.
Irac-Astaud, M., and Rideau, G. (1992). On the existence of quantum bihamiltonian systems: The harmonic oscillator case, Universitรฉ Paris VII preprint, PAR-LPTM92.
Irac-Astaud, M., and Rideau, G. (1993). Letters in Mathematical Physics, 29, 197.
Irac-Astaud, M., and Rideau, G. (1994). Theorical and Mathematical Physics, 99, 658.
Jannussis, A. (1993). Journal of Physics A, 26, L233.
Jannussis, A., Brodimas, G., and Mignani, R. (1991). Journal of Physics A, 24, L775.
Jordan, T. F., Mukunda, N., and Pepper, S. V. (1963). Journal of Mathematical Physics, 4, 1089.
Katriel, J., and Quesne, C. (1996). Journal of Mathematical Physics, 37, 1650.
Khare, A. (1992). Journal of Physics A, 25, L749.
Khare, A. (1993). Journal of Mathematical Physics, 34, 1277.
Khare, A., Mishra, A. K., and Rajasekaran, G. (1993a). International Journal of Modern Physics A, 8, 1245.
Khare, A., Mishra, A. K., and Rajasekaran, G. (1993b). Modern Physics Letters A, 8, 107.
Kosiลski, P., Majewski, M., and Maลlanka, P. (1997). Journal of Physics A, 30, 3983.
Kuryshkin, V. (1980). Annales de la Fondation Louis de Broglie, 5, 111.
McDermott, R. J., and Solomon, A. I. (1994). Journal of Physics A, 27, L15.
Macfarlane, A. J. (1989). Journal of Physics A, 22, 4581.
Macfarlane, A. J. (1994). Journal of Mathematical Physics, 35, 1054.
Manโko, V. I., Marmo, G., Sudarshan, E. C. G., and Zaccaria, F. (1997). Physica Scripta, 55, 528.
Meljanac, S., and Milekoviฤ, M. (1996). International Journal of Modern Physics A, 11, 1391.
Meljanac, S., Milekoviฤ, M., and Pallua, S. (1994). Physics Letters B, 328, 55.
Mishra, A. K., and Rajasekaran, G. (1991a). Pramana - Journal of Physics, 36, 537.
Mishra, A. K., and Rajasekaran, G. (1991b). Pramana - Journal of Physics, 37, 455(E).
Ohnuki, Y., and Kamefuchi, S. (1982). Quantum Field Theory and Parastatistics, Springer-Verlag, Berlin.
Paolucci, A., and Tsohantjis, I. (1997). Physics Letters A, 234, 27.
Plyushchay, M. S. (1996a). Modern Physics Letters A, 11, 397.
Plyushchay, M. S. (1996b). Annals of Physics (N.Y.), 245, 339.
Polychronakos, A. P. (1992). Physical Review Letters, 69, 703.
Quesne, C. (1994a). Journal of Physics A, 27, 5919.
Quesne, C. (1994b). Physics Letters A, 193, 245.
Quesne, C. (1995). Modern Physics Letters A, 10, 1323.
Quesne, C., and Vansteenkiste, N. (1995). Journal of Physics A, 28, 7019.
Quesne, C., and Vansteenkiste, N. (1996). Helvetica Physica Acta, 69, 141.
Quesne, C., and Vansteenkiste, N. (1997). Czechoslovak Journal of Physics, 47, 115.
Quesne, C., and Vansteenkiste, N. (1998). Physics Letters A, 240, 21.
Quesne, C., and Vansteenkiste, N. (1999). Helvetica Physica Acta, 72, 71.
Rideau, G. (1992). Letters in Mathematical Physics, 24, 147.
Rubakov, V. A., and Spiridonov, V. P. (1988). Modern Physics Letters A, 3, 1337.
Solomon, A. I. (1998). In Fifth International Conference on Squeezed States and Uncertainty Relations, Balatonfured, Hungary, May 27โ31, 1997, D. Han, J. Janszky, Y. S. Kim, and V. I. Manโko, eds., NASA Conference Publication 1998-206855, NASA Goddard Space Flight Center, Greenbelt, Maryland, p. 157.
Sukhatme, U. P., Rasinariu, C., and Khare, A. (1997). Physics Letters A, 234, 401.
Sun, C.-P., and Fu, H.-C. (1989). Journal of Physics A, 22, L983.
Tsohantjis, I., Paolucci, A., and Jarvis, P. D. (1997). Journal of Physics A, 30, 4075.
Vasiliev, M. A. (1991). International Journal of Modern Physics A, 6, 1115.
Veselov, A. P., and Shabat, A. B. (1993). Funktsionalnyi Analiz i Ego Prilozheniia, 27, 1.
Witten, E. (1981). Nuclear Physics B, 185, 513.
Wolfes, J. (1974). Journal of Mathematical Physics, 15, 1420. |
warning/0003/hep-ph0003132.html | ar5iv | text | # Mesonic Anapole Form Factors of the Nucleons
## 1 Introduction
Experimental determination of the strangeness form factors of the nucleons aims at the coupling of the $`Z^0`$ boson to the strange quarks and measures the interference term between the $`\gamma `$ and $`Z^0`$ exchange interactions between electrons and protons in electron-nucleons scattering . If the nucleons have non-zero anapole moments, these contribute an axial current component to the electron-nucleon coupling, which has to be taken into account as a radiative correction in the extraction of the strangeness form factors from the measured asymmetry .
Anapole moments are induced by parity violating meson fluctuations of the nucleons, as e.g. $`W^\pm `$ fluctuations and $`\pi `$-loops with one parity violating (PV) vertex . For the nucleons the anapole form factors may be defined as $`F_A^{p,n}(0)=F_A^S(q^2)\pm F_A^V(q^2)`$, where $`F_A^S(q^2)`$ and $`F_A^V(q^2)`$ are the isoscalar and isovector form factors in the current matrix element
$$<p^{}|j_\mu ^A(0)|p>=ie\overline{u}(p^{})\frac{F_A^S(q^2)+\tau _3F_A^V(q^2)}{m_N^2}[q^2\gamma _\mu \gamma _52im_Nq_\mu \gamma _5]u(p).$$
$`(1.1)`$
Here $`m_N`$ is the nucleon mass, and $`q`$ is the momentum transfer to the nucleon ($`q=p^{}p)`$. The space favoring metric $`q^2=\stackrel{}{q}^2q_0^2`$ will be employed throughout.
The anapole moment due to pionic fluctuations of the proton was estimated in ref. to be $`0.12f_\pi `$, where $`f_\pi `$ is the PV pion-nucleon coupling constant, the standard value of which is $`4.510^7`$ . An anapole moment this small does not affect the empirical extraction of the strangeness magnetic moment of the proton in PV electron scattering. A significant effect from the anapole moment would require that it be larger by at least an order of magnitude.
A calculation of the anapole moment of the proton based on the chiral quark model is reported here. As the pseudoscalar and vector mesons couple directly to constituent quarks in this model, the anapole moment arises as a sum of contributions from PV meson loop fluctuations of the constituent quarks and PV meson exchange currents along with PV โpolarization currentsโ induced by the PV meson exchange interaction between quarks. Current conservation demands the presence of all of these combined. The calculation is therefore reminiscent of the calculation of nuclear anapole moments in refs. . This provides an alternate approach to the baryonic loop calculations, which recently have been recast into the form of chiral perturbation theory . The calculation based on the chiral quark model allows a unified treatment of baryon structure and the baryon spectrum based on the same Hamiltonian. It takes all baryonic intermediate states into account, as the constituent quarks lack excited states. A major difference between the quark model and effective baryon field theory approaches is that while the $`SU(6)`$ aspect of the baryon wave function plays plays a major role in the former it plays none whatever in the latter. The magnetic moments of the baryons provide an example of a set of observables for which the 3-quark structure of the baryon wave function plays the leading role, and meson loops give but small contributions in the quark model.
In the chiral quark model the magnitude of the calculated anapole moment is determined by the PV meson-nucleon coupling constants. By taking into account both the pion and vector meson exchange contributions, the value for the anapole moment of the proton is found to be $`0.910^8`$, with a very large uncertainty margin, that mostly is set by the poorly known values of the PV meson-nucleon coupling constants. The pion contributions alone give rise to a positive value for the anapole moment ($`+1.8810^8`$), in agreement with results that are obtained in chiral perturbation theory . With the โrecommendedโ values for the PV meson-nucleon coupling constants there are strong cancellations between the pion and vector meson contributions to the anapole moment. The pionic contribution to the anapole moment is found to be somewhat smaller in the chiral quark model than in the pion loop calculation at the baryon level because of a tendency to cancellation between the pion loop and pion exchange current contributions.
This paper falls into 7 sections. In sections 2 and 3 the pion loop and pion exchange and polarization current contributions respectively are calculated. In sections 4 and 5 the corresponding vector meson loop contributions are derived. In section 6 the $`\rho \pi `$ and $`\omega \pi `$ loop contributions are calculated. Section 7 contains a discussions of the results and their implications.
## 2 Pion loop contributions to the anapole form factors
The parity violating (PV) and parity conserving (PC) couplings to constituent quarks have the expressions:
$$_{\pi qq}^{PV}=g_{\pi qq}^W\overline{\psi }(\stackrel{}{\tau }\times \stackrel{}{\varphi })_3\psi ,$$
$`(2.1a)`$
$$_{\pi qq}^{PC}=i\frac{f_{\pi qq}}{m_\pi }\overline{\psi }\gamma _5\gamma _\mu _\mu \stackrel{}{\varphi }\stackrel{}{\tau }\psi .$$
$`(2.1b)`$
Here $`\psi `$ represents the field of the constituent $`u`$ and $`d`$ quarks and $`\stackrel{}{\varphi }`$ the pion field. By means of standard quark model algebra the pion-quark coupling constants may be expressed in terms of the corresponding pion-nucleon coupling constants as
$$g_{\pi qq}^W=g_{\pi NN}^W=\frac{f_\pi }{\sqrt{2}},$$
$`(2.2a)`$
$$f_{\pi qq}^{PC}=\frac{3}{5}f_{\pi NN}0.6.$$
$`(2.2b)`$
Here $`f_\pi `$ is the PV $`\pi N`$ coupling in the notation of ref. and $`f_{\pi NN}`$ is the pseudovector $`\pi N`$-coupling.
The PV pion fluctuations of the constituent quarks, which contribute to the anapole moments of the constituent quarks are illustrated by the Feynman diagrams in Fig. 1. For the evaluation of these loop amplitudes the pion and quark current operators are e
$$j_\mu ^\pi =e(_\mu \stackrel{}{\varphi }\times \stackrel{}{\varphi })_3,$$
$`(2.3a)`$
$$j_\mu ^q=ie\overline{\psi }(\frac{1}{6}+\frac{\tau _3}{2})\gamma _\mu \psi .$$
$`(2.3b)`$
Here it has been assumed that the anomalous Pauli terms in the e.m. current of the constituent quarks are unimportant .
Minimal substitution of the e.m. field $`\stackrel{}{A}`$ in the chiral coupling (2.1b) generates a contact coupling term:
$$_{\pi \gamma qq}=ie\frac{f_{\pi qq}}{m_\pi }\overline{\psi }\gamma _5\gamma _\mu A_\mu (\stackrel{}{\varphi }\times \stackrel{}{\tau })_3\psi ,$$
$`(2.4)`$
which generates contact coupling diagrams in addition to the pion loop diagrams in Fig. 1. For e.m. couplings, the amplitudes that are obtained with the PV coupling (2.1b) and the contact coupling (2.4) are equivalent to those obtained with the pseudoscalar coupling
$$_{\pi qq}=ig_{\pi qq}\overline{\psi }\gamma _5\stackrel{}{\varphi }\stackrel{}{\tau }\psi ,$$
$`(2.5)`$
if $`g_{\pi qq}=(2m_q/m_\pi )f_{\pi qq}`$ ($`m`$ is the constituent quark mass).
The contributions from the pion loop fluctuations in Figs. 1 to the e.m. current of the $`u[d]`$ quarks may be expressed as
$$j_\mu =[+]2eg_{\pi qq}^Wg_{\pi qq}\frac{d^4p}{(2\pi )^4}\frac{(k_b+k_a)_\mu }{(k_b^2+m_\pi ^2)(k_a^2+m_\pi ^2)}\{\frac{1}{\gamma pim}\gamma _5\gamma _5\frac{1}{\gamma pim}\}$$
$$+2e\frac{1}{3}[\frac{2}{3}]g_{\pi qq}^Wg_{\pi qq}\frac{d^4k}{(2\pi )^4}\frac{1}{k^2+m_\pi ^2}\{\frac{1}{\gamma p_bim}+\gamma _\mu \frac{1}{\gamma p_aim}\gamma _5$$
$$\gamma _5\frac{1}{\gamma p_0im}\gamma _\mu \frac{1}{\gamma pim}\}.$$
$`(2.6)`$
Here $`k_b=p_{out}p,k_a=p_{in}p`$ in the first integral (Fig. 1a) and $`p_b=p_{out}k,p_a=p_{in}k`$ in the second integral (Fig. 1b). The overall coefficients are those for $`u`$-quarks, whereas the overall coefficients for the $`d`$-quarks are given in the square brackets \[โฆ\]. The amplitudes (2.6) contain the appropriate flavor factors (Table 1).
The current operator (2.6) satisfies the current conservation constant $`q_\mu j_\mu =0`$ only by addition of the contributions to the current from the self-energy diagrams in Fig. 2. These divergent terms do not contribute to the $`q`$-dependence of the form factor. The calculation of the anapole moment therefore should be carried out by dropping these terms and enforcing current conservation on the loop amplitudes (2.6) after subtraction of the divergent unphysical pole terms. The current transversality requirement may be imposed directly by replacement of the operators $`\gamma _\mu `$ and $`(k_a+k_b)_\mu `$ in (2.6) by
$$\gamma _\mu \gamma _\mu \frac{\gamma q}{q^2}q_\mu $$
$`(2.7a)`$
$$k_{a_\mu }+k_{b_\mu }k_{a_\mu }+k_{b_\mu }\frac{q(k_a+k_b)}{q^2}q_\mu .$$
$`(2.7b)`$
This procedure is equivalent to projecting out the longitudinal component from the final result.
Pions are assumed to decouple from constituent quarks at the chiral symmetry restoration scale $`\mathrm{\Lambda }_\chi 4\pi f_\pi 1.2`$ GeV. The loop integrals should accordingly be cut off at or about that momentum scale. This should be done so as to maintain the transversality condition $`q_\mu j_\mu =0`$. If in the quark coupling (second) term in (2.5) the opinion propagator is replaced by
$$\frac{1}{m_\pi ^2+k^2}v(k^2)=\frac{1}{m_\pi ^2+k^2}(\frac{\mathrm{\Lambda }^2m_\pi ^2}{\mathrm{\Lambda }^2+k^2})^2,$$
$`(2.8)`$
the transversality condition is maintained, provided that the product of the pion propagators in the pion coupling (first) term in (2.5) is replaced by :
$$\frac{1}{k_b^2+m_\pi ^2}\frac{1}{k_a^2+m_\pi ^2}\frac{v(k_b^2)v(k_a^2)}{k_a^2k_b^2}.$$
$`(2.9)`$
Once the anapole moments $`F_A^u`$ and $`F_A^d`$ of the $`u`$ and $`d`$-quarks respectively have been calculated from the expression (2.6), the corresponding proton- and neutron anapole form factors are obtained by the standard quark model expressions as
$$\frac{F_A^p}{m_N^2}=\frac{4}{3}\frac{F_A^u}{m^2}\frac{1}{3}\frac{F_A^d}{m^2},$$
$`(2.10a)`$
$$\frac{F_A^n}{m_N^2}=\frac{4}{3}\frac{F_A^d}{m^2}\frac{1}{3}\frac{F_A^u}{m^2}.$$
$`(2.10b)`$
The expressions for the anapole form factors of the proton and the neutron are then found to be
$$F_A^p(q^2)=\frac{m_N^2}{m^2}\{F_\pi (q^2)+\frac{2}{3}F_q(q^2)\},$$
$`(2.11a)`$
$$F_A^n(q^2)=\frac{m_N^2}{m^2}\{F_\pi (q^2)+\frac{7}{3}F_q(q^2).$$
$`(2.11b)`$
Here $`F_\pi `$ and $`F_q`$ are the contributions to the anapole moment of the $`u`$-quark from the pion and quark coupling loops (Figs. 1 a,b) in (2.6):
$$F_\pi (q^2)=\frac{m^2}{q^2}\{\frac{g_{\pi qq}^Wg_{\pi qq}}{4\pi ^2}_0^1dxx_0^1dy\{(log\frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_\pi ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\pi ^2}{H_2(\mathrm{\Lambda }_\chi ^2)})$$
$$(\mathrm{})_{q^2=0}\},$$
$`(2.12a)`$
$$F_q(q^2)=\frac{m^2}{q^2}\frac{g_{\pi qq}^Wg_{\pi qq}}{12\pi ^2}_0^1๐x(1x)_0^1๐y$$
$$\{([m^2(1x^2)q^2(1x)^2y(1y)]K_1(q^2)+log\frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_\pi ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\pi ^2}{H_1(\mathrm{\Lambda }_\chi ^2)})$$
$$(\mathrm{})_{q^2=0}\}.$$
$`(2.12b)`$
Here the notation $`(\mathrm{})_{q^2=0}`$ indicates that the value of the preceeding bracket at $`q^2=0`$ should be subtracted from it. This subtraction takes into account the self energy terms.
The auxiliary functions in (2.12) are defined as
$$H_1(M^2)=M^2x+m^2(1x)^2+q^2(1x)^2y(1y),$$
$`(2.13a)`$
$$H_2(M^2)=M^2x+m^2(1x)+q^2x^2y(1y).$$
$`(2.13b)`$
Finally the function $`K_1(q^2)`$ is defined as
$$K_1(q^2)=\frac{1}{H_1(m_\pi ^2)}\frac{1}{H_1(\mathrm{\Lambda }_\chi ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\pi ^2}{H_1^2(\mathrm{\Lambda }_\chi ^2)}.$$
$`(2.14)`$
The calculated pion loop contributions to the anapole form factors of the proton and the neutron are shown in Fig. 3 as functions of momentum transfer. In the calculation the โstandardโ value for $`f_\pi `$ (2.2a) $`4.510^7`$ was used. The constituent quark mass was set to $`m=340`$ MeV and the value of the cut-off $`\mathrm{\Lambda }_\chi `$ was taken to be 1.2 GeV.
The numerical values for pion loop contributions to the anapole moments of the proton and the neutron were found to be $`F_A^p(0)=3.2210^8`$ and $`F_A^n(0)=1.4510^8`$. These values should be added to those obtained from the pion exchange and polarization currents calculated below. Note similarity between the calculated proton and neutron anapole moments, which indicates that main component of the pion loop contribution to the anapole moment is the isoscalar term, in agreement with other findings . The expression for the pion loop contribution to the isoscalar combination of the anapole form factors here is formally similar to that, which appears in the derivation of the loop contribution at the baryon level with only nucleon intermediate states , the only difference being the replacement of quark masses and coupling constants by the corresponding nucleon masses and coupling constants. The expressions for the isovector amplitudes is different however, which is natural, as the quark level calculation takes into account all baryonic intermediate states, and in particular the $`\mathrm{\Delta }_{33}`$ resonance.
## 3 Pion exchange and polarization currents
a. Parity violating pion exchange currents
The pion exchange current operator, which is generated by the pion-quark couplings (2.1), (2.4) and the e.m. currents of the pions and constituent quarks (2.3a) are illustrated diagrammatically in Fig. 4. In the absence of vertex form factors the expressions for the parity violating contact and pion current exchange current operators are
$$j_\mu ^\pi (C)=\frac{ieg_{\pi qq}^Wg_{\pi qq}}{2m}[\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2]\{\frac{\gamma _5^1\gamma _\mu ^1}{k_2^2+m_\pi ^2}+\frac{\gamma _5^2\gamma _\mu ^2}{k_1^2+m_\pi ^2}\},$$
$`(3.1a)`$
$$j_\mu ^\pi (\pi )=eg_{\pi qq}^Wg_{\pi qq}[\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2]\frac{(\gamma _5^2\gamma _5^1)(k_1k_2)_\mu }{(k_1^2+m_\pi ^2)(k_2^2+m_\pi ^2)}.$$
$`(3.1b)`$
Here the Dirac equation has been invoked for the quarks. The 4-momentum fractions imparted to the two quarks are denoted $`k_1,k_2`$ respectively, so that $`q=k_1+k_2`$.
The current conservation condition on the PV pion exchange current $`j_\mu ^{ex}(C)+j_\mu ^{ex}(\pi )`$ is :
$$q_\mu j_\mu ^{ex}=[V_\pi ^{PV}(p_1^{},p_2^{},p_1+q,p_2)j_0^1(p_1+q,p_1)$$
$$j_0^1(p_1^{},p_1q)V_\pi ^{PV}(p_1^{}q,p_2^{},p_1,p_2)]$$
$$+(12).$$
$`(3.2)`$
Here $`j_0^1`$ is the charge density operator of a single quark and $`V_\pi ^{PV}`$ is the parity violating pion exchange potential for constituent quarks. The initial and final 4-momenta of the quark pair are denoted $`p_1,p_2`$ and $`p_1^{},p_2^{}`$ respectively. The expression for the PV $`\pi `$-exchange interaction is
$$V_\pi ^{PV}=ig_{\pi qq}g_{\pi qq}^W(\stackrel{}{\tau }^1\times \stackrel{}{\tau }^2)_3\frac{\gamma _5^1\gamma _5^2}{k^2+m_\pi ^2}.$$
$`(3.3)`$
A practical consequence of the continuity equation (3.2) is the requirement that the exchange current and exchange induced polarization current contributions combine to the proper non-singular form (1.1).
The pion exchange Yukawa functions in the exchange current operators (3.1) as well as in the potential (3.3) should be cut-off at the chiral restoration scale as were the pion loops. Current conservation is maintained if in the contact current operator (3.1a) and the potential the Yukawa function $`1/(k^2+m_\pi ^2)`$ is modified as in (2.8), and the product of pion Yukawa functions in the pionic current (3.1b) is replaced as in (2.9) .
The current conservation condition (3.2) ensures that the exchange current operator compensates for the non-conservation of the single quark e.m. current in the nucleon states, in the presence of a parity violating pion exchange interaction between the constituent quarks. If the negative parity component of the proton wave function is neglected, transversality has to be enforced on the net exchange current contribution in the calculation of the anapole moment. This procedure will be implemented here, although the polarization current that is induced by the negative parity component is also considered explicitly below.
For the calculation of the matrix elements of the pion exchange current operator it is conducive to rewrite it in the spin representation. To lowest order in the inverse quark masses the operators (3.1) combine to the expression
$$\stackrel{}{j}^\pi =e\frac{g_{\pi qq}^Wg_{\pi qq}}{2m}(\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2)\{\frac{\stackrel{}{\sigma }^2}{k_1^2+m_\pi ^2}+\frac{\stackrel{}{\sigma }^1}{k_2^2+m_\pi ^2}$$
$$\frac{(\stackrel{}{\sigma }^1\stackrel{}{k}_1\stackrel{}{\sigma }^2\stackrel{}{k}_2)(\stackrel{}{k}_1\stackrel{}{k}_2)}{(k_1^2+m_\pi ^2)(k_2^2+m_\pi ^2)}\}.$$
$`(3.4)`$
In addition to the parameters that appear in the calculation of the pion loop contributions to the anapole moment the exchange current contribution also depends on the proton wave function and thus on the confinement scale. The orbital part of proton wave function will here be described by the oscillator function (Table 2)
$$\psi (\stackrel{}{r},\stackrel{}{\rho })=(\frac{\omega }{\sqrt{\pi }})^3e^{(r^2+\rho ^2)\omega ^2/2},$$
$`(3.5)`$
which appears translationally invariant quark models as well as in the covariant integrable quark model for the baryons developed in ref. . Here $`\stackrel{}{r}`$ and $`\stackrel{}{\rho }`$ are Jacobi coordinates for the 3-quark system. With a constant flavor-spin hyperfine interaction the baryon spectrum is well described with the value $`\omega =311`$ MeV , whereas in recent a meson exchange model for the spectrum $`\omega =1240`$ MeV . The latter value is most consistent with the present meson exchange model for the anapole moment.
The orbital matrix element of the pion exchange contact current (3.1a) may be written as
$$j_\mu (C)=\frac{eg_{\pi qq}^Wg_{\pi qq}}{2m}\frac{m_\pi }{4\pi }(\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2)[i\gamma _5^1\gamma _\mu ^1+i\gamma _5^2\gamma _\mu ^2]M_\rho (q)M_r(q),$$
$`(3.6)`$
where $`M_\rho (q)`$ and $`M_r(q)`$ are the orbital matrix elements:
$$M_\rho (q)=4\pi (\frac{\omega }{\sqrt{\pi }})^3_0^{\mathrm{}}๐\rho \rho ^2j_0(\frac{q\rho }{\sqrt{6}})e^{\rho ^2\omega ^2},$$
$`(3.7)`$
$$M_r(q)=4\pi (\frac{\omega }{\sqrt{\pi }})^3_0^{\mathrm{}}๐rr^2j_0(\frac{qr}{\sqrt{2}})y_0(m_\pi r\sqrt{2})e^{r^2\omega ^2}.$$
$`(3.8)`$
In the latter integral $`y_0(x)`$ is a cut-off Yukawa function defined as
$$y_0(m_\pi r\sqrt{2})=\frac{e^{m_\pi r\sqrt{2}}}{m_\pi r\sqrt{2}}(\frac{\mathrm{\Lambda }}{m_\pi })\frac{e^{\mathrm{\Lambda }_\chi r\sqrt{2}}}{\mathrm{\Lambda }_\chi r\sqrt{2}}\frac{\mathrm{\Lambda }_\chi ^2m_\pi ^2}{2\mathrm{\Lambda }_\chi m_\pi }e^{\mathrm{\Lambda }_\chi r\sqrt{2}}.$$
$`(3.9)`$
For proton and neutron states with spin $`z`$-component $`+1/2`$, the matrix element of the spin flavor operator calculated with the wave functions listed in Table 3 is
$$<(\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2)(i\gamma _5^1\gamma _3^2+i\gamma _5^2\gamma _3^2)><(\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2)(\sigma _3^1+\sigma _3^2)>=\frac{4}{3}.$$
$`(3.10)`$
The contribution of pion exchange contact current to the anapole moment of the proton and the neutron then takes the form
$$F_{A,C}^p(q^2)=F_{A,C}^n(q^2)=\frac{2}{3}\frac{m_N^2}{q^2}\frac{g_{\pi qq}^Wg_{\pi qq}}{4\pi }\frac{m_\pi }{m}[M_\rho (q)M_r(q)M_\rho (0)M_r(0)].$$
$`(3.11)`$
The subtraction of the matrix element at $`q=0`$ is required, unless a consistent calculation of the polarization current is performed, in which case the matrix element at $`q=0`$ is cancelled by a corresponding term in the latter. This point will be treated in detail below.
The calculation of the matrix element of the pionic exchange current operator (3.1b) is somewhat more complicated. The result for this contribution to the anapole moment may be cast into the form
$$F_{A,\pi }^p(q^2)=F_{A,\pi }^n(q^2)=\frac{m_N^2}{q^2}\frac{g_{\pi qq}^Wg_{\pi qq}}{\pi }\frac{1}{18m}M_\rho (q)_0^1dx\{(m_\pi ^{}(x)$$
$$4\pi (\frac{\omega }{\sqrt{\pi }})^3_0^{\mathrm{}}drr^2e^{r^2\omega ^2}\{j_0(\xi )[2y_0(m_\pi ^{}r\sqrt{2})y_1(m_\pi ^{}r\sqrt{2})]$$
$$j_2(\xi )[y_0(m_\pi ^{}r\sqrt{2})+y_1(m_\pi ^{}r\sqrt{2})]\}$$
$$(\mathrm{})_{q^2=0}\}.$$
$`(3.12)`$
Here the variable $`\xi `$ is defined as
$$\xi =qr(\frac{1}{2}x)\sqrt{2},$$
$`(3.13)`$
and the function $`y_1(m_\pi ^{}r\sqrt{2})`$ is defined as:
$$y_1(m_\pi ^{}r\sqrt{2})=e^{m_\pi ^{}r\sqrt{2}}(\frac{\mathrm{\Lambda }^{}}{m_\pi ^{}})e^{\mathrm{\Lambda }^{}r\sqrt{2}}$$
$$+\frac{1}{2}\frac{\mathrm{\Lambda }^2m_\pi ^2}{m_\pi ^{}\mathrm{\Lambda }^{}}(1\mathrm{\Lambda }^{}r\sqrt{2})e^{\mathrm{\Lambda }^{}r\sqrt{2}}.$$
$`(3.14)`$
The mass parameters $`m_\pi ^{}`$ and $`\mathrm{\Lambda }^{}`$ are functions of the integration variable $`x`$:
$$m_\pi ^{}(x)=\sqrt{m_\pi ^2+q^2x(1x)},$$
$`(3.15a)`$
$$\mathrm{\Lambda }^{}(x)=\sqrt{\mathrm{\Lambda }^2+q^2x(1x)}.$$
$`(3.15b)`$
The calculated pion exchange current contributions to the anapole form factors of the proton and the neutron are shown in Fig. 5 The exchange current contributions are shown both as calculated with the oscillator parameters $`\omega =311`$ MeV and 1240 MeV in order to obtain an estimate of the theoretical uncertainty of the calculated values. The calculated value of the pion exchange current contribution to the anapole moment of the proton (and the neutron) is -1.80 $`10^8`$ with $`\omega =`$ 311 MeV and $`1.1510^8`$ with $`\omega `$ =1240 MeV. As pointed out above the larger oscillator parameter is consistent with the meson exchange model for the baryon wave functions in ref. . While the wave function dependence of the anapole moment contribution itself is thus not very strong the exchange current contribution to the anapole form factor is very strong.
For both parameter values a tendency for cancellation between the exchange current and the pion loop contributions to the anapole moment is visible. This then suggests that the net values of the anapole moments of the nucleons will be small and only of the order $`10^8`$. In the present quark model calculation the exchange current contribution to the anapole moment is purely isoscalar, but the pion loop contribution is about 2 times larger in the case of the proton than in the case of the neutron. This result differs from that found in ref. , where the PV pion loop fluctuations of the proton were considered. The isoscalar nature in that loop calculation is a consequence of dominance of the contributions of the terms, in which the e.m. coupling is to the intermediate pions over the contributions from the terms in which the e.m. coupling is to the intermediate nucleon.
b. Parity violating pion exchange induced polarization current
The parity violating pion exchange interaction (3.3) induces a negative parity component into the proton wave function. In view of the small magnitude of the PV pion exchange potential, this component may be calculated by means of first order perturbation theory as
$$\psi ^{}=\underset{n}{}\frac{<n|V_\pi ^{PV}|0>}{E_nE_0}\psi _n^{},$$
$`(3.16)`$
where $`\psi _n^{}`$ are the wave functions that describe the negative parity excitations of the nucleons, and $`E_nE_0`$ are the corresponding resonance excitation energies.
The electromagnetic current operator of the constituent quarks:
$$\stackrel{}{j}_q=e(\frac{1}{6}+\frac{\tau _3}{2})\{\frac{\stackrel{}{p}^{}+\stackrel{}{p}}{2m}+\frac{i}{2m}\stackrel{}{\sigma }\times \stackrel{}{q}\},$$
$`(3.17)`$
will have non-vanishing matrix elements between the positive $`(\psi ^+)`$ and negative $`(\psi ^{})`$ parity components of the nucleon wave function, which contribute to the anapole form factor of the nucleon. These contributions are referred to as polarization currents . The polarization current then takes the form
$$\stackrel{}{j}^{pol}=\underset{n}{}(\psi ^+,\underset{q}{}\stackrel{}{j}_q\frac{<n|V_\pi ^{PV}|0>}{E_nE_0}\psi _n^{})$$
$$\underset{n}{}(\psi _n^{}\frac{<0|V_\pi ^{PV}|n>}{E_nE_0},\underset{q}{}\stackrel{}{j}_q\psi ^+).$$
$`(3.18)`$
This expression is illustrated schematically in Fig. 7.
The only $`\frac{1}{2}^{}`$ nucleon states in the $`P`$-shell of the baryons are the $`N(1535)`$ and $`N(1650)`$ resonances. The former is a member of an $`S=1/2`$ negative parity doublet and the latter a member of an $`S=\frac{3}{2}`$ negative parity quartet, in the usual assignment . With this assignment only the $`N(1535)`$ resonance contributes to the sum (3.18). As the $`P`$-shell lies $`600`$ MeV above the nucleon, it is justified to truncate the sum in (3.18) after the $`P`$-shell states, in view of the much larger energy denominators of the resonances in the $`F`$ and higher shells.
The required matrix element of the PV pion exchange interaction may be evaluated using the oscillator wave functions of the covariant quark model in ref. , which are listed in Table 2. The only non-vanishing matrix elements are
$$<N(1535)^+,\frac{1}{2}|V_\pi ^{PV}|N(939)^+,\frac{1}{2}>=i\frac{\sqrt{2}}{3}\frac{g_{\pi qq}g_{\pi qq}^W}{4\pi }\frac{m_\pi ^2}{m},$$
$`(3.19)`$
where $``$ is the radial matrix element
$$=\sqrt{2}\omega (\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}๐rr^3\overline{y}_1(m_\pi r\sqrt{2})e^{r^2\omega ^2}.$$
$`(3.20)`$
In (3.19) the $``$ sign applies for protons, and the $`+`$ sign for neutrons. The Yukawa function $`\overline{y}_1`$ is defined as
$$\overline{y}_1(m_\pi r\sqrt{2})=(1+1/m_\pi r\sqrt{2})\frac{e^{m_\pi r\sqrt{2}}}{m_\pi r\sqrt{2}}(\frac{\mathrm{\Lambda }_\chi }{m_\pi })^2(1+1/\mathrm{\Lambda }_\chi r\sqrt{2})\frac{e^{\mathrm{\Lambda }_\chi r\sqrt{2}}}{\mathrm{\Lambda }_\chi r\sqrt{2}}$$
$$\frac{1}{2}((\frac{\mathrm{\Lambda }_\chi }{m_\pi }^21)e^{\mathrm{\Lambda }_\chi r\sqrt{2}}.$$
$`(3.21)`$
Evaluation of the matrix element (3.18) of the convection current part (3.17) of the single quark current operator yields the following contribution to the anapole form factors of the proton and the neutron:
$$F_{A,conv}^p=F_{A,conv}^n=\eta \frac{m_N^2}{q^2}\frac{1}{3}\frac{g_{\pi qq}g_{\pi qq}^W}{4\pi }\frac{m_\pi ^2}{m^2}\frac{\omega }{\mathrm{\Delta }}(\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}๐\rho \rho ^2e^{\rho ^2\omega ^2}[j_0(\sqrt{\frac{2}{3}}q\rho )1].$$
$`(3.22)`$
Here $`\mathrm{\Delta }`$ is the energy dominator $`\mathrm{\Delta }=1535939=596`$ MeV. Note that this is an isoscalar term, as required by current conservation with the isoscalar pion exchange current operator considered above. The factor $`\eta `$ is a correction factor that has to be chosen so that the continuity equation that links the exchange current and polarization current corrections is satisfied. If all negative parity states in the sum over $`n`$ in (3.18) are taken into account, this factor would be unity. In order that no spurious pole at $`q^2=0`$ occur in the sum of pion exchange current and polarization current matrix elements appear in the present perturbative calculation of the latter, the value for the correction factor $`\eta `$ has to be chosen as
$$\eta =\frac{2}{3}\frac{m\mathrm{\Delta }}{\omega ^2}.$$
$`(3.23)`$
If the oscillator parameter $`\omega `$ is taken to be $`\omega `$ = 311 MeV , the numerical value for $`\eta `$ is 1.4, which indicates that truncating the sum over $`n`$ in (3.18) is fairly good approximation. In the case of the larger value 1240 MeV for $`\omega `$ the expression (3.23) gives $`\eta `$ = 0.09, which indicates that higher lying excited states contribute significantly.
Once the correction factor $`\eta `$ is included in the expression (3.22), there is in principle no need to subtract the values of the matrix elements at $`q=0`$ from the exchange (3.11), (3.12) and polarization current contributions (3.22) to the anapole moment. Above the value of the matrix elements at $`q=0`$ have been subtracted explicitly, as it gives an explicit indication that the pole term at $`q^2`$ = 0 has to cancel out.
The spin current component of the single quark current operator (3.17) gives rise to the following contribution to the anapole moment of the proton and the neutron:
$$F_{A,spin}^{p,n}=\eta (1,\frac{1}{3})\frac{2}{27}\frac{m_N^2}{m^2}\frac{m_\pi ^2}{\mathrm{\Delta }}\frac{g_{\pi qq}g_{\pi qq}^W}{4\pi }\omega (\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}๐\rho \rho ^4e^{\rho ^2\omega ^2}[j_0(\sqrt{\frac{2}{3}}q\rho )+j_2(\sqrt{\frac{2}{3}}q\rho )].$$
$`(3.24)`$
Here the term 1 in the first bracket on the r.h.s. applies in the case of the proton and the term $`1/3`$ in the case of the neutron.
The pion exchange induced polarization current contributions to the anapole form factors of the neutron and the proton have been plotted along with the corresponding pion exchange current contributions in Figs. 5 and 6. The polarization current contribution to the anapole moment of the nucleon depends strongly on the wave function model. With the wave function parameter value $`\omega `$ = 311 MeV the polarization current contribution to the anapole moment of the proton is $`3.3810^8`$ and with the parameter value $`\omega `$ = 1240 MeV it is $`0.1910^8`$. The corresponding contributions to the anapole moment of the neutron are $`0.3810^8`$ and $`0.0210^8`$ respectively.
The net combination of pion exchange and polarization current contributions to the anapole form factors of the proton and the neutron are shown in Fig. 8. The net calculated pionic contribution to the anapole moment of the proton is โ including the pion loop contribution - is $`1.8810^8`$ and that to the anapole moment of the neutron is $`0.2810^8`$ (for $`\omega `$ = 1240 MeV).
## 4 Vector meson loop contributions to the anapole moment
### 4.1 $`\rho `$-meson loop contributions
The parity violating $`\rho `$-nucleon coupling is by convention written in terms of 3 different flavor coupling terms
$$=i\overline{\psi }_N\{h_\rho ^0\stackrel{}{\tau }\stackrel{}{\rho }_\mu +h_\rho ^1\rho _\mu ^3+h_\rho ^2\frac{3\tau _3\rho _\mu ^3\stackrel{}{\tau }\stackrel{}{\rho }_\mu }{2\sqrt{6}}\}\gamma _\mu \gamma _5\psi _N.$$
$`(4.1)`$
Here $`\psi _N`$ is the nucleon and $`\stackrel{}{\rho }_\mu `$ the isovector field of the $`\rho `$-meson. The coupling constants $`h_\rho `$ have been determined phenomenologically only within wide uncertainly ranges, the โrecommendedโ values being $`h_\rho ^0=30g_w,h_\rho ^1=0.5g_w`$ and $`h_\rho ^2=25g_w(g_w=3.810^8)`$ . Standard quark model algebra then implies that the PV coupling of $`\rho `$-meson to constituent quarks be
$$=i\overline{\psi }\{h_{\rho qq}^0\stackrel{}{\tau }\stackrel{}{\rho }_\mu +h_{\rho qq}^1\rho _\mu ^3+h_{\rho qq}^2\frac{3\tau _3\rho _\mu ^3\stackrel{}{\tau }\stackrel{}{\rho }_\mu }{2\sqrt{6}}\}\gamma _\mu \gamma _5\psi ,$$
$`(4.2)`$
where $`\psi `$ represents the quark fields. The PV $`\rho `$-quark coupling constants $`h_{\rho qq}`$ are determined by the corresponding $`\rho `$-nucleon coupling constants as
$$h_{\rho qq}^0=\frac{3}{5}h_\rho ^0,h_{\rho qq}^1=h_\rho ^1,h_{\rho qq}^2=\frac{3}{5}h_\rho ^2.$$
$`(4.3)`$
The PV $`\rho `$-meson loop contributions to anapole moments of the constituent $`u`$ and $`d`$ quarks are illustrated diagrammatically in Fig. 9. The flavor factors for the different flavor coupling terms $`h_\rho `$ are listed in Table 1 for the different diagrams. The calculation of these loop contributions require the $`\rho `$-quark coupling Lagrangian
$$_{\rho qq}=ig_{\rho qq}\overline{\psi }\gamma _\mu \stackrel{}{\tau }\stackrel{}{\rho }_\mu \psi ,$$
$`(4.4)`$
where $`g_{\rho qq}=g_{\rho NN}2.6`$ and the e.m. current operator for the $`\rho `$-meson
$$j_\mu =\pm ie\{\rho _\nu ^{}_\mu \rho _\nu \rho _\nu ^{}_\nu \rho _\mu \}.$$
$`(4.5)`$
With these couplings and the flavor factors listed in Table 1 the expressions for the $`\rho `$-meson loop amplitudes, for $`u`$ quarks, where e.m. coupling is to the internal quark (Fig. 9a), takes the form
$$j_\mu (u)=\frac{1}{3}(2h_{\rho qq}^1+\frac{h_{\rho qq}^2}{\sqrt{6}})eg_{\rho qq}\frac{d^4k}{(2\pi )^4}\frac{\delta _{\alpha \beta }+\frac{k_\alpha k_\beta }{m_p^2}}{k^2+m_p^2}$$
$$\{\gamma _\alpha \gamma _5\frac{1}{\gamma p_bim}\gamma _\mu \frac{1}{\gamma p_aim}\gamma _\beta \gamma _5$$
$$+\gamma _\alpha \frac{1}{\gamma p_bim}\gamma _\mu \frac{1}{\gamma p_aim}\gamma _\beta \gamma _5\}.$$
$`(4.6)`$
Here the notation is the same as in eqn. (2.6). The same expression applies for $`d`$-quarks provided that the flavor factor $`(2h_{\rho qq}^1+h_{\rho qq}^2/\sqrt{6})/3`$ with the PV $`\rho `$-quark couplings is replaced by the corresponding factor $`(h_{\rho qq}^0+h_{\rho qq}^1/3h_{\rho qq}^2/\sqrt{6}).`$
The expression for the contribution of the corresponding $`\rho `$-meson loop to fluctuations to the $`u`$-quark current, illustrated in Fig. 9b, where the e.m. coupling is to the $`\rho `$-meson is
$$j_\mu (u)=2(h_{\rho qq}^0\frac{1}{\sqrt{6}}h_{\rho qq}^2)eg_{\rho qq}\frac{d^4p}{(2\pi )^4}\frac{1}{k_b^2+m_p^2}\frac{1}{k_a^2+m_p^2}$$
$$\{(k_a+k_b)_\mu (\delta _{\alpha \beta }+\frac{k_{a\alpha }k_{a\beta }}{m_\rho ^2})(\delta _{\beta \delta }+\frac{k_{b\beta }k_{b\delta }}{m_\rho ^2})$$
$$\{\gamma _\delta \gamma _5\frac{1}{\gamma pim}\gamma _\alpha +\gamma _\delta \frac{1}{\gamma pim}\gamma _\alpha \gamma _5\}$$
$$\gamma _\delta \gamma _5(\delta _{\beta \delta }+\frac{k_{b\delta }k_{b\beta }}{m_\rho ^2})\frac{k_{a\beta }}{\gamma pim}(\delta _{\mu \alpha }+\frac{k_{a\mu }k_{a\alpha }}{m_\rho ^2})\gamma _\alpha $$
$$\gamma _\delta (\delta _{\beta \delta }+\frac{k_{b\delta }k_{b\beta }}{m_\rho ^2})\frac{k_{a\beta }}{\gamma pim}(\delta _{\mu \alpha }+\frac{k_{a\mu }k_{a\alpha }}{m_\rho ^2})\gamma _a\gamma _5$$
$$\gamma _\alpha \gamma _5(\delta _{\mu \alpha }+\frac{k_{b\mu }k_{b\alpha }}{m_\rho ^2})\frac{k_{b\beta }}{\gamma pim}(\delta _{\beta \delta }+\frac{k_{a\delta }k_{a\beta }}{m_\rho ^2})\gamma _\delta $$
$$\gamma _\alpha (\delta _{\mu \alpha }+\frac{k_{b\mu }k_{b\alpha }}{m_\rho ^2})\frac{k_{b\beta }}{\gamma pim}(\delta _\beta \delta +\frac{k_{a\delta }k_{a\beta }}{m_\rho ^2})\gamma _\delta \gamma _5\}.$$
$`(4.7)`$
The corresponding operator for $`\rho `$-meson fluctuations of the $`d`$-quark in this case is obtained simply by change of the overall sign. The sum of the current operators (4.6) and (4.7) satisfy the tranversality condition when added to the combination of $`\rho `$-meson self energy diagrams (cf. Fig. 2). As those only add a $`q`$-independent infinite contribution, which has to be subracted, they are dropped here out and the transversality condition on the $`\rho `$-meson loop currents (4.1) and (4.7) is enforced as in the case of the pion loop diagrams above.
The current operators (4.6) and (4.7) simplify significantly by leaving out the terms that are inversely proportional to $`m_\rho ^2`$ in the numerator of the $`\rho `$-meson propagators. As those terms have been found to be of very small numerical significance for $`q^2<m_p^2`$ in other calculations of related type , they are dropped here, especially as their effect is small in comparison to that of the wide uncertainty in the PV $`\rho `$-meson coupling constants .
The contributions to the anapole moments of the $`u`$ and $`d`$ quarks from the $`\rho `$-meson loops with the e.m. coupling to the internal quark line (Fig. 9 a) and to the $`\rho `$-meson (Fig. 9 b), respectively, are then
$$F_{A,q}^{u,d}(q^2)=f^{u,d}\frac{m^2}{q^2}\frac{g_{\rho qq}}{4\pi ^2}_0^1๐x(1x)_0^1๐y$$
$$\{[m^2(1x)^2q^2xq^2(1x)^2y(1y)]\overline{K}_1(q^2,m_\rho ^2)$$
$$+log\frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_\rho ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{H_1^2(m_\rho ^2)}\}\{\mathrm{}\}_{q^2=0},$$
$`(4.8a)`$
$$F_{A,\rho }^{u,d}(q^2)=g^{u,d}\frac{m^2}{q^2}\frac{g_{\rho qq}}{4\pi ^2}_0^1๐xx_0^1๐y$$
$$\{[2m^2x(1x)+xq^2+2x^2q^2y(1y)]\overline{K}_2(q^2,m_\rho ^2)$$
$$+6log\frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_\rho ^2)}6x\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{H_2^2(m_\rho ^2)}\}\{\mathrm{}\}_{q^2=0}.$$
$`(4.8b)`$
Here $`f^{u,d}`$ and $`g^{u,d}`$ are the flavor factors
$$f^u=\frac{2h_{\rho qq}^1}{3}+\frac{h_{\rho qq}^2}{\sqrt{6}},f^d=h_{\rho qq}^0+\frac{h_{\rho qq}^1}{3}\frac{h_{\rho qq}^2}{\sqrt{6}},$$
$`(4.9a)`$
$$g^u=h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}},g^d=h_{\rho qq}^0+\frac{h_{\rho qq}^2}{2\sqrt{6}}.$$
$`(4.9b)`$
The functions $`K_1(q^2)`$ and $`K_2(q^2)`$ are defined as (cf. ref. )
$$\overline{K}_1(q^2,m_\rho ^2)=\frac{1}{H_1(m_\rho ^2)}\frac{1}{H_1(\mathrm{\Lambda }_\chi ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{H_1^2(\mathrm{\Lambda }_\chi ^2)},$$
$`(4.10a)`$
$$\overline{K}_2(q^2,m_\rho ^2)=\frac{1}{H_2(m_\pi ^2)}\frac{1}{H_2(\mathrm{\Lambda }_\chi ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{H_2^2(\mathrm{\Lambda }_\chi ^2)}.$$
$`(4.10b)`$
The $`\rho `$โmeson loop contributions to the anapole moments of the proton and the neutron may then be calculated using eqs. (2.10). In Fig. 10 the calculated anapole form factors of the proton and the neutron are shown. The magnitudes are the similar to those given by the corresponding pion loop contributions, but the signs are opposite. In the numerical calculations the โrecommendedโ values for the PV $`\rho `$โnucleon coupling constants were employed. The calculated values of the $`\rho `$โmeson loop contributions to the anapole moments of the proton and the neutron were found to be $`0.4410^8`$ and $`1.5210^8`$ respectively.
The $`\rho `$meson loops here are those, which are induced by the PV $`\rho `$meson-constituent quark couplings, which in turn are implied by the $`\rho `$nucleon couplings (4.1). In addition to these, there may also appear a PV $`\rho \rho \gamma `$ coupling, akin to the anapole moment term in the e.m. current operator. That coupling has been considered in ref. and would generate further PV loop contributions to the constituent quark current. These are conserved by the form of the coupling, and inversely proportional to $`m_\rho ^2`$. In view of the uncertain value of the PV $`\rho \rho \gamma `$ coupling strength and as the terms of order $`m_\rho ^2`$ were dropped from the loop calculation above, we have for consistency not included these loop contributions in the present calculation.
### 4.2 $`\omega `$-meson loop contributions
The standard expression for the parity violating $`\omega `$-nucleon coupling is
$$=i\overline{\psi }_N\{h_\omega ^0\omega _\mu +h_\omega ^1\tau ^3\omega _\mu \}\gamma _\mu \gamma _5\psi _N.$$
$`(4.11)`$
The โrecommended valuesโ for the PV $`\omega `$-nucleon coupling constants are $`h_\omega ^0=5g_w`$ and $`h_\omega ^1=3g_w`$, with a large uncertainty margin .
The corresponding PV coupling of $`\omega `$ mesons to constituent quarks is
$$=i\overline{\psi }\{h_{\omega qq}^0\omega _\mu +h_{\omega qq}^1\tau ^3\omega _\mu \}\gamma _\mu \gamma _5\psi .$$
$`(4.12)`$
The $`SU(6)`$ quark model for the nucleon wave functions leads to the following relations between the PV $`\omega `$-nucleon and the corresponding PV $`\omega `$-quark coupling constants:
$$h_{\omega qq}^0=\frac{h_\omega ^0}{3},h_{\omega qq}^1=\frac{3}{5}h_\omega ^1.$$
$`(4.13)`$
We shall rely on these relations here.
For the calculation of the contributions from the PV $`\omega `$-meson loop fluctuations of the constituent $`u`$ and $`d`$ quarks to their anapole moments, the $`\omega `$-quark coupling
$$=ig_{\omega qq}\overline{\psi }\gamma _\mu \omega _\mu \psi $$
$`(4.14)`$
is employed. The $`\omega `$-quark coupling constant determined from the corresponding $`\omega `$-nucleon vector coupling constant $`g_{\omega NN}`$ is $`g_{\omega qq}=g_{\omega NN}/3`$ by the quark model. The recent Nijmegen model for the nucleon-nucleon interaction determines $`g_{\omega NN}=10.35`$, from which it follows that $`g_{\omega qq}=3.45`$. This value is used here.
Because of the neutrality of the $`\omega `$ meson the e.m. field only couples to the internal quark in the PV $`\omega `$ meson loop contributions. The expressions for these contributions to the anapole moments of the $`u`$ and $`d`$ quarks, then take the form
$$F_{A,\omega }^{u,d}(q^2)=\frac{m^2}{q^2}(h_{\omega qq}^0h_{\omega qq}^1)_0^1๐x(1x)_0^1๐y$$
$$\{[m^2(1x^2)q^2xq^2(1x)^2y(1y)]\overline{K}_1(q^2,m_\omega ^2)$$
$$+log\frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_\omega ^2)}x\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{H_1^2(m_\rho )}\}\{\mathrm{}\}_{q^2=0}.$$
$`(4.15)`$
The expressions for the $`\omega `$-loop contributions to the nucleon anapole form factors are then given by the equations (2.10). In Fig. 10 the numerical values for the $`\omega `$-loop contributions to the anapole form factors of the proton and the neutron are also shown. The numerical values for the $`\omega `$-loop contributions to the anapole moments of the nucleons are found to be โ0.097$`10^8`$ (proton) and 0.34$`10^8`$ (neutron) (Table 4). The $`\omega `$ meson exchange current contributions of to the anapole moments are expected to be very small in view of the neutrality of the $`\omega `$meson and the small values of the PV $`\omega `$ coupling constants as compares to the corresponding $`\rho `$meson couplings..
## 5 Vector meson exchange and polarization currents
a. Parity violating $`\rho `$ meson exchange currents
The PV $`\rho `$-meson exchange current operators, which are implied by the $`\rho `$-quark couplings (4.12), (4.4) are illustrated diagrammatically in Fig. 11. The contact current is customarily derived as a pair current operator , but may also be derived as a contact current as in the case of the pion exchange current $`j_\mu (C)`$ (3.1a) above. To see this one may rewrite the the parity conserving $`\rho `$-quark coupling (4.4) as
$$_{\rho qq}=g_{\rho qq}\overline{\psi }((p^{}+p)_\mu \frac{1}{2m}\sigma _{\mu \nu }_\nu )\stackrel{}{\rho }_\mu \stackrel{}{\tau }\psi .$$
$`(5.1)`$
Minimal substitution of the e.m. vector potential $`A_\mu `$ in this coupling yields the contact coupling
$$_{\gamma \rho qq}=e\frac{g_{\rho qq}}{2m}\overline{\psi }\{\sigma _{\mu \nu }A_\nu (\stackrel{}{\rho }_\mu \times \stackrel{}{\tau })_3A_\mu (\stackrel{}{\rho }_\mu \stackrel{}{\tau }+\rho _{\mu 3}\}\psi $$
$`(5.2)`$
in analogy with (2.4) in the case of the pion.
The expressions for the $`\rho `$-meson contact and $`\rho `$-current exchange current operators are found to be
$$j_\mu ^\rho (C)=ie\frac{g_{\rho qq}}{2m}\frac{1}{k_2^2+m_\rho ^2}\{(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})(\stackrel{}{\tau }^1\times \stackrel{}{\tau }^2)_3\gamma _\nu ^2\gamma _5^2\sigma _{\nu \mu }^1$$
$$[(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\stackrel{}{\tau }^1\stackrel{}{\tau }^2+(h_{\rho qq}^0\frac{h_{\rho qq}^2}{\sqrt{6}})\tau _3^2+h_{\rho qq}^1(1+\tau _3^1)+\frac{3}{2\sqrt{6}}\tau _3^1\tau _3^2)]\}+(12),$$
$`(5.3a)`$
$$j_\mu ^\rho (\rho )=ieg_{\rho qq}(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\frac{(\stackrel{}{\tau }^1\times \stackrel{}{\tau }^2)_3}{(k_1^2+m_\rho ^2)(k_2^2+m_\rho ^2)}$$
$$\{(k_1k_2)_\mu [\gamma _\nu ^2\gamma _5^2\gamma _\nu ^1+\gamma _\nu ^1\gamma _5^1\gamma _\nu ^2]$$
$$\gamma _\nu ^2\gamma _5^2k_{1\nu }\gamma _\mu ^1+\gamma _\nu ^1\gamma _5^1k_{2\nu }\gamma _\mu ^2+\gamma _\mu ^2\gamma _5^2\gamma _\nu ^1k_{2\nu }\gamma _\mu ^1\gamma _5^1\gamma _\nu ^2k_{1\nu }\}.$$
$`(5.3b)`$
In the spin representation the sum of the local parts of the $`\rho `$-meson exchange current operators (5.3), to lowest order in $`1/m^2`$, reduce to the expression
$$\stackrel{}{j}^\rho =e(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\frac{g_{\rho qq}}{2m}(\stackrel{}{\tau }^1\times \stackrel{}{\tau }^2)_3$$
$$\{(\stackrel{}{\sigma }^1\times \stackrel{}{\sigma }^2)\{\frac{1}{k_1^2+m_\rho ^2}+\frac{1}{k_2^2+m_\rho ^2}\}\frac{1}{(k_1^2+m_\rho ^2)(k_2^2+m_\rho ^2)}$$
$$\{(\stackrel{}{\sigma }^1\times \stackrel{}{\sigma }^2)(\stackrel{}{k}_1\stackrel{}{k}_2)(\stackrel{}{k}_1\stackrel{}{k}_2)+\stackrel{}{\sigma }^2\stackrel{}{k}_1\stackrel{}{\sigma }^1\times \stackrel{}{k}_1\sigma ^1k_2\stackrel{}{\sigma }^2\times \stackrel{}{k}_2\stackrel{}{\sigma }^2\stackrel{}{\sigma }^1\stackrel{}{k}_1\times \stackrel{}{k}_2+\stackrel{}{\sigma }^1\stackrel{}{\sigma }^2\stackrel{}{k}_1\times \stackrel{}{k}_2\}\}$$
$$+e\frac{g_{\rho qq}}{2m}\{\frac{\stackrel{}{\sigma }^2}{k_2^2+m_\rho ^2}[(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\stackrel{}{\tau }^1\stackrel{}{\tau }^2+(h_{\rho qq}^0\frac{h_{\rho qq}^2}{\sqrt{6}})\tau _3^2+h_{\rho qq}^1(1+\tau _3^1)+\frac{3}{2\sqrt{6}}\tau _3^1\tau _3^2)]$$
$$+\frac{\stackrel{}{\sigma }^1}{k_1^2+m_\rho ^2}[(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\stackrel{}{\tau }^1\stackrel{}{\tau }^2+(h_{\rho qq}^0\frac{h_{\rho qq}^2}{\sqrt{6}})\tau _3^1+h_{\rho qq}^1(1+\tau _3^2)+\frac{3}{2\sqrt{6}}\tau _3^1\tau _3^2)].$$
$`(5.4)`$
Apart from the coupling constants, this PV $`\rho `$-meson exchange current operator has a formal similarity to that of the corresponding PV pion exchange current operator (3.4). In contrast to the latter the $`\rho `$-meson exchange current operator is an isovector operator.
Comparison of the expressions (3.4) and (5.4) for the PV $`\pi `$ and $`\rho `$ meson exchange current operator makes it possible to derive the $`\rho `$-meson exchange contribution to the anapole moments of the proton and the neutron by analogy. The required spin-isospin matrix elements are given in Table 3.
The contribution of the seagull current (5.3a) to the anapole moment of the proton and the neutron becomes becomes
$$F_{A,C}(q)=4\frac{m_N^2}{q^2}h_C(p,n)\frac{g_{\rho qq}}{4\pi }$$
$$\frac{m_\rho }{m}[M_\rho (q)\overline{M}_r(q)M_\rho (0)\overline{M}_r(0)].$$
$`(5.5)`$
Here the coefficients $`h_C(p,n)`$ are defined as
$$h_C(p,n)=\pm (h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\frac{1}{2}(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\frac{5}{12}(h_{\rho qq}^0\frac{h_{\rho qq}^2}{\sqrt{6}})\frac{h_{\rho qq}^1}{4}+\frac{1}{12}h_{\rho qq}^1\frac{h_{\rho qq}^2}{8\sqrt{6}}.$$
$`(5.6)`$
The upper and lower signs in this expression applies to protons and neutrons respectively. The function $`M_\rho (0)`$ is defined in (3.7), and the matrix element $`\overline{M}_r(q)`$ is defined as
$$\overline{M}_r(q)=4\pi (\frac{\omega }{\sqrt{\pi }})^3_0^{\mathrm{}}๐rr^2j_0(\frac{qr}{\sqrt{2}})y_0(m_\rho \sqrt{2}r)e^{r^2\omega ^2}$$
$`(5.7)`$
The cut-off Yukawa function $`y_0`$ is defined in (3.9), the $`\rho `$-meson mass here having been substituted for the pion mass.
The contribution to the anapole moment from the PV $`\rho `$-meson exchange current operator (5.3b) (second term in (5.4)), where the e.m. coupling is to the $`\rho `$-meson field, takes the form (cf. (3.12))
$$F_{A,\rho }^p(q)=F_{A,n}^n(q)=\frac{m_N^2}{q^2}(h_{\rho qq}^0\frac{h_{\rho qq}^2}{2\sqrt{6}})\frac{g_{\rho qq}}{4\pi }\frac{1}{m}M_\rho (q)$$
$$_0^1dx\{m_\rho ^{}(x)4\pi (\frac{\omega }{\sqrt{\pi }})^3_0^{\mathrm{}}drr^2e^{r^2\omega ^2}$$
$$\{2j_0(\xi )[2y_0(m_\rho ^{}r\sqrt{2})(1\frac{\stackrel{}{q}^2}{8m_\rho ^2})y_1(m_\rho ^{}r\sqrt{2})]$$
$$j_2(\xi )[y_0(m_\rho ^{}r\sqrt{2})+y_1(m_\rho ^{}r\sqrt{2})](\mathrm{})_{q=0}\}.$$
$`(5.8)`$
Here the notation is the same as used in the expression (3.11). The mass function $`m_\rho ^{}(x)`$ is defined as (cf. 3.15a)
$$m_\rho ^{}(x)=\sqrt{m_\rho ^2+q^2x(1x)}.$$
$`(5.9)`$
The net calculated $`\rho `$-meson exchange current contributions $`F_A^\rho (q,C)+F_A^\rho (q,\rho )`$ to the anapole form factors of the proton and the neutron are shown in Figs. 12 and 13 respectively for two different values of the wave function parameter $`\omega `$. With $`\omega `$ = 311 MeV these contributions to the anapole moments of the proton and the neutron are $`1.1810^8`$ and $`0.8510^8`$ respectively. With the larger value $`\omega =1240`$ MeV, which is consistent with the Hamiltonian model for the baryon spectrum in ref. , the corresponding values are $`1.7910^8`$ and $`0.7310^8`$ respectively.
In ref. another type of PV $`\rho `$meson exchange current was considered. That exchange current arises from the PV $`\rho \rho \gamma `$ coupling, which corresponds to the anapole coupling of the $`\gamma `$ to the nucleons. This isoscalar exchange current is conserved by construction, and inversely proportional to $`m_\rho ^2`$. Because of the uncertain value of the PV $`\rho \rho \gamma `$ coupling strength and as the terms of order $`m_\rho ^2`$ have not been included in the calculation above, this isoscalar $`\rho `$ meson exchange current has not been included here for reasons of consistency.
b. Parity violating $`\rho `$ exchange induced polarization current
The parity violating $`\rho `$ meson exchange interaction between constituent quarks contributes to the negative parity component of the proton wave function. This interaction takes the form
$$V_\rho ^{PV}=\frac{g_{\rho qq}}{4\pi }\{h_{\rho qq}^0\stackrel{}{\tau }^1\stackrel{}{\tau }^2+\frac{1}{2}h_{\rho qq}^1(\tau _3^1+\tau _3^2)$$
$$+\frac{1}{2\sqrt{6}}h_{\rho qq}^2(3\tau _3^1\tau _3^2\stackrel{}{\tau }^1\stackrel{}{\tau }^2)\}$$
$$\{(\stackrel{}{\sigma }^1\stackrel{}{\sigma }^2)\{\frac{\stackrel{}{p}_1\stackrel{}{p}_2}{2m},f(r)\}_+\}+i\stackrel{}{\sigma }^1\times \stackrel{}{\sigma }^2[\frac{\stackrel{}{p}_1\stackrel{}{p}_2}{2m},f(r)]\}$$
$$+\frac{g_{\rho qq}}{4\pi }\frac{h_{\rho qq}^1}{2}(\tau _3^1\tau _3^2)(\stackrel{}{\sigma }^1+\stackrel{}{\sigma }^2)[\frac{\stackrel{}{p}_1\stackrel{}{p}_2}{2m},f(r)].$$
$`(5.10)`$
The radial Yukawa function $`f(r)`$ is defined as
$$f(r)=\frac{e^{m_\rho r}}{r}\frac{e^{\mathrm{\Lambda }_\chi r}}{r}\frac{m_\rho ^2}{2\mathrm{\Lambda }_\chi }(\frac{\mathrm{\Lambda }_\chi ^2}{m_\rho ^2}1)e^{\mathrm{\Lambda }_\chi r}.$$
$`(5.11)`$
Here $`r=|\stackrel{}{r}_1\stackrel{}{r}_2|`$. The $`\rho `$-exchange induced PV polarization current is then given by the general expression (3.18) with $`V_\rho ^{PV}`$ substituted in place of $`V_\pi ^{PV}`$. In order to be consistent with the approximate treatment of the $`\rho `$meson exchange current operators above, where only the local part of the operator was retained, the $`\rho `$meson exchange induced polarization contribution is calculated here using only the local (commutator) terms in the $`\rho `$meson exchange interaction.
If the sum over negative parity states is approximated by the contributions from the lowest lying negative parity resonances, $`N(1535)`$ and $`\mathrm{\Delta }(1620)`$, the only matrix elements that are required are
$$<N(1535)^+,\frac{1}{2}|V_\rho ^{PV}|N(939)^+,\frac{1}{2}>=i\sqrt{2}\frac{g_{\rho qq}}{4\pi }\frac{m_\rho ^2}{m}(\frac{h_{\rho qq}^0}{2}\frac{h_{\rho qq}^1}{3})_\rho ,$$
$`(5.12a)`$
$$<\mathrm{\Delta }(1620)^+,\frac{1}{2}|V_\rho ^{PV}|N(939)^+,\frac{1}{2}>=2i\sqrt{2}\frac{g_{\rho qq}}{4\pi }\frac{m_\rho ^2}{m}(\pm \frac{h_{\rho qq}^1}{6}+\frac{h_{\rho qq}^2}{2\sqrt{6}})_\rho ^a.$$
$`(5.12b)`$
Here the orbital matrix element $`_\rho `$ is defined as
$$_\rho =\sqrt{2}\omega (\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}๐rr^3e^{r^2\omega ^2}\overline{y}_1(m_\rho r\sqrt{2}),$$
$`(5.13)`$
Here the wave function models in Table 2 has again been employed. The function $`\overline{y}_1`$ is defined as in (3.12).
Evaluation of the matrix element (3.18) of the convection current part of the single quark current (3.17), with $`V_\pi ^{PV}`$ replaced by $`V_\rho ^{PV}`$, with the sum over $`n`$ truncated to include only the contributions from the $`N(1535)`$ and $`\mathrm{\Delta }(1620)`$ negative parity resonances then leads to the following anapole form factor contributions to the proton and the neutron:
$$F_{A,conv}^p=F_{A,conv}^n=\eta \frac{m_N^2}{q^2}\frac{g_{\rho qq}}{4\pi }\frac{m_\rho ^2}{m^2}\{\frac{1}{\mathrm{\Delta }_1}(\frac{h_{\rho qq}^0}{2}\frac{1}{3}h_{\rho qq}^1)_\rho $$
$$+\frac{2}{\mathrm{\Delta }_2}((\frac{h_{\rho qq}^1}{6}+\frac{h_{\rho qq}^2}{2\sqrt{6}})_\rho \}\omega (\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}d\rho \rho ^2e^{\rho ^2\omega ^2}[j_0(\sqrt{\frac{2}{3}}q\rho )1].$$
$`(5.14)`$
Here $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$ are the mass differences $`m[N(1535)]m_N=596`$ MeV and $`m[\mathrm{\Delta }(1620)m_N]=681`$ MeV respectively. The same renormalization correction factor $`\eta `$ (3.23) as was employed in the expression for the pion exchange induced parity violating polarization current has been used here. The required renormalization constant was derived explicitly from the continuity equation in the case of pion exchange. It cannot be calculated in the same way in the case of $`\rho `$-meson exchange, because the PV $`\rho `$-coupling (4.1) does not satisfy the required transversality condition.
The spin current component of the single quark current operator (3.17) gives the following $`\rho `$-meson exchange induced polarization current contribution to the anapole form factors of the proton and the neutron:
$$F_{A,spin}^{p,n}=\eta (1,\frac{1}{3})\frac{2}{9}\frac{g_{\rho qq}}{4\pi }\frac{m_N^2m_\rho ^2}{m^2}\{\frac{1}{\mathrm{\Delta }_1}(\frac{h_{\rho qq}^0}{2}\frac{1}{3}h_{\rho qq}^1)_\rho $$
$$+\frac{2}{3\mathrm{\Delta }_2}((\pm \frac{h_{\rho qq}^1}{6}+\frac{h_{\rho qq}^2}{2\sqrt{6}})_\rho \}$$
$$\omega (\frac{\omega }{\sqrt{\pi }})^34\pi _0^{\mathrm{}}๐\rho \rho ^4e^{\rho ^2\omega ^2}[j_0(\sqrt{\frac{2}{3}}q\rho )+j_2(\sqrt{\frac{2}{3}}q\rho )].$$
$`(5.15)`$
Here the term 1 in the bracket $`(1,1/3)`$ on the r.h.s. applies in the case of the proton and the term $`1/3`$ in the case of the neutron.
The $`\rho `$meson exchange induced contributions to the anapole form factor of the proton is shown in Fig. 12 as obtained with two different values for the wave function parameter $`\omega `$. The corresponding contributions to the anapole form factor of the neutron are shown in Fig. 13. For $`\omega =`$ 1240 MeV the $`\rho `$-exchange induced polarization current contribution to the anapole moment of the proton is $`+0.08810^8`$ and that of the neutron is $`0.04710^8`$ (Table 4).
## 6 Meson loops with transition couplings
In addition to the pion and vector meson loop contributions to the anapole form factors, loop contributions with e.m. pseudoscalar-vector meson transition couplings also contribute to the anapole form factors. These are illustrated by the Feynman diagrams in Fig. 14.
The $`\rho \pi \gamma `$ transition coupling may be descibed by the current matrix element
$$<\pi ^a(k^{})|j_\mu (0)|\rho _\nu ^b(k)>=i\frac{g_{\rho \pi \gamma }}{m_\rho }ฯต_{\mu \nu \lambda \sigma }k_\lambda k_\sigma ^{}\delta ^{ab}.$$
$`(6.1)`$
The $`\rho \pi \gamma `$ coupling constant may be calculated from the known radiative widths of the $`\rho `$-mesons: $`\mathrm{\Gamma }(\rho ^\pm \pi ^\pm \gamma )=68`$ keV, $`\mathrm{\Gamma }(\rho ^0\pi ^0\gamma )=119`$ keV. These yield $`g_{\rho ^\pm \pi ^\pm \gamma }=0.57`$ and $`g_{\rho ^0\pi ^0\gamma }=0.75`$.
The corresponding transition current matrix element for the $`\omega \pi ^0\gamma `$ transition is
$$<\pi ^0(k^{})|j_\mu (0)|\omega _\nu (k)>=i\frac{g_{\omega \pi \gamma }}{m_\omega }ฯต_{\mu \nu \lambda \sigma }k_\lambda k_\sigma ^{}.$$
$`(6.2)`$
From the radiative width of the $`\omega `$-meson: $`\mathrm{\Gamma }(\omega \pi ^0\gamma )=717`$ MeV one obtains $`g_{\omega \pi \gamma }=1.84`$.
The current operators that describe the $`\rho \pi \gamma `$ transition loops on $`u`$\- and $`d`$-quarks in Fig. 9, where one of the hadronic vertices is parity violating takes the form
$$j_\mu ^{u,d}=i\frac{G^{u,d}}{m_\rho }\frac{d^up}{(2\pi )^4}\{\frac{ฯต_{\mu \nu \lambda \sigma }k_{a\lambda }k_{b\sigma }}{(k_b^2+m_\pi ^2)(k_a^2+m_\rho ^2)}\frac{1}{\gamma pim}\gamma _\nu $$
$$+\gamma _\nu \frac{1}{\gamma pim}\frac{ฯต_{\mu \nu \lambda \sigma }k_{a\lambda }k_{b\sigma }}{(k_b^2+m_\rho ^2)(k_a^2+m_\pi ^2)}\}.$$
$`(6.3)`$
Here the coupling constant combinations $`G^u`$ and $`G^d`$ have been defined as
$$G^{u,d}=2g_{\pi qq}^Wg_{\rho qq}g_{\rho ^\pm \pi ^\pm \gamma }+g_{\pi qq}[2h_{\rho qq}^0g_{\rho ^\pm \pi ^\pm \gamma }$$
$$+(h_{\rho qq}^0\pm h_{\rho qq}^1+\frac{h_{\rho qq}^2}{\sqrt{6}})g_{\rho ^0\pi ^0\gamma }].$$
$`(6.4)`$
The only difference between $`G^u`$ and $`G^d`$ is in the sign of the $`h_{\rho qq}^1`$ term, which is positive in the case of $`G^4`$ and negative in the case of $`G^d`$.
Because of the presence of the Levi-Civita symbol in the current expressions (6.3), the apparent ultraviolet divergences drop out. Similarly the terms $`k_\alpha k_\beta /m_\rho ^2`$ in the vector meson propagators drop out. In line with the cutting off of the loop integrals of the pion and vector meson loop contributions above at the chiral symmetry restoration scale, a similar cutoff is applied here by insertion of a form factor,
$$F(k_\pi ^2,k_\rho ^2)=\frac{\mathrm{\Lambda }_\chi ^2m_\pi ^2}{\mathrm{\Lambda }_\chi ^2+k_\pi ^2}\frac{\mathrm{\Lambda }_\chi ^2m_\rho ^2}{\mathrm{\Lambda }_\chi ^2+k_\rho ^2},$$
$`(6.5)`$
in the integrands, where $`k_\pi `$ and $`k_\rho `$ denote the 4-momenta of the $`\pi `$ and $`\rho `$ mesons.
To reduce expressions (6.3) to standard form (1.1), it is convenient to employ the relation
$$ฯต_{\mu \nu \lambda \sigma }(p^{}+p)_\lambda q_\sigma \gamma _\nu =q^2\gamma _\mu \gamma _52im\gamma _5q_\mu ,$$
$`(6.6)`$
where $`q_\mu =p_\mu ^{}p_\mu `$. the PV $`\rho \pi `$ loop contributions to the anapole moments of the $`u`$ and $`d`$ quarks then become
$$F_{A,\rho \pi }^{u,d}(q^2)=\frac{G^{u,d}}{16\pi ^2}\frac{m}{m_\rho }_0^1๐xx_0^1๐y(1x)[1+x(12y)]$$
$$\{\frac{1}{G(m_\rho ,m_\pi )}\frac{1}{G(m_\rho ,\mathrm{\Lambda }_\chi )}\frac{1}{G(\mathrm{\Lambda }_\chi ,m_\pi )}+\frac{1}{G(\mathrm{\Lambda }_\chi ,\mathrm{\Lambda }_\chi )}\}.$$
$`(6.6)`$
Here the function $`G(m_1,m_2)`$ has been defined as
$$G(m_1,m_2)=m^2(1x)^2+m_1^2x(1y)+m_2^2xy+q^2x^2y(1y).$$
$`(6.7)`$
The contribution from the PV loops that involve the $`\omega \pi \gamma `$ transition in the meson line may be calculated by the same method. This contribution obtains no contribution from the PV $`\pi `$-quark coupling (2.1a), which only contributes to loop amplitudes with charged pions. The PV $`\pi \gamma `$ transition loops leads to the following anapole form factors of the $`u`$ and $`d`$ quarks:
$$F_{A,\omega \pi }^{u,d}(q^2)=\frac{H^{u,d}}{16\pi ^2}\frac{m}{m_\omega }_0^1๐xx_0^1๐y(1x)[1+x(12y)]$$
$$\{\frac{1}{G(m_\omega ,m_\pi )}\frac{1}{G(m_\omega ,\mathrm{\Lambda }_\chi )}\frac{1}{G(\mathrm{\Lambda }_\chi ,m_\omega )}+\frac{1}{G(\mathrm{\Lambda }_\chi ,\mathrm{\Lambda }_\chi )}\}.$$
$`(6.8)`$
Here a cut-off factor of the form (6.5) with $`m_\rho `$ replaced by $`m_\omega `$ has been included in the expression. The factors $`H^{u,d}`$ represent the coupling constant combinations
$$H^{u,d}=g_{\pi qq}(h_{\omega qq}^0\pm h_{\omega qq}^1)g_{\omega \pi \gamma }.$$
$`(6.9)`$
The $`\rho \pi `$ and $`\omega \pi `$ loop contributions to the anapole moment of the proton, as calculated using the expressions (2.10) have been plotted in Fig. 15. The contribution from the $`\rho \pi `$ loops is the larger one because of the larger PV $`\rho `$quark coupling strengths. These contribute $`0.6710^8`$ and $`0.6510^8`$ to the anapole moments of the proton and the neutron (Table 4). The corresponding contributions from the $`\omega \pi `$ loops to the anapole moments of the proton and the neutron are $`0.1210^8`$ and $`0.03510^8`$ respectively.
The $`\rho \pi \gamma `$ and $`\omega \pi \gamma `$ transition couplings also give rise to parity violating exchange current operators. These are illustrated diagrammatically in Fig. 16. The corresponding 2-quark currents may be derived by using the meson-nucleon couplings in sections 2 and 4 and the $`\rho \pi \gamma `$ and $`\omega \pi \gamma `$ transition couplings (6.1) and (6.2).
These exchange current operators do not however contribute to the anapole form factors of the nucleon. The exchange current operator that involves the PV pion-quark coupling (2.1a) is spin-independent, and vanishes in nucleon states because of the antisymmetric Levi-Civita symbol in the e.m. pion-vector meson transition couplings (6.1) and (6.2). The exchange currents that involve the PV vector meson couplings (4.2) (4.11) have no local component, which would contribute to the anapole form factor.
## 7 Discussion
The calculated net anapole form factors of the proton and the neutron are shown in Fig. 17. These results take into account all the loop and exchange current contributions calculated above. As a satisfactory description of the nucleon radii with the constituent quark model of ref. requires that a form factor be associated with the constituent quarks , the calculated anapole form factors have been multiplied by the corresponding form factor in Fig. 17. The form of the constituent quark form factor used was $`exp(r^2q^2/6)`$, where the mean square radius value is $`r^2=0.133`$ fm<sup>2</sup> . The calculated anapole moments are also shown without this quark form factor modification.
The corresponding contributions to the anapole moments of the proton and the neutron are listed in Table 4. The calculated value of the anapole moment of the proton is $`0.9010^8`$, the magnitude of whichwhich is considerably smaller than what is obtained by considering only the pionic contributions. The main difference from the PV pion-nucleon loop estimate in ref. and the chiral perturbation theory estimate in ref. is the appearance of exchange and polarization current contributions at the quark level. The present calculation suggests that the net mesonic anapole moment is far too small to lead to any notable correction to the extraction of the strangeness magnetic moment in the SAMPLE experiment . The net mesonic contribution to the anapole moment of the neutron is $`0.6810^8`$.
The calculated anapole moment has a wide $`100\%`$ theoretical uncertainty, which mainly arises from the large uncertainty in the PV meson-nucleon couplings , which are used as input parameters in the calculation. The loop amplitudes have a cut-off sensitivity, but this is fairly small, once the cut-off is taken to be of the order of the chiral symmetry restoration scale. The exchange current and polarization current contributions have an additional dependence on the baryon wave function model, which to some extent is limited by requirement of consistency with the calculated baryon spectrum .
A main qualitative result of the present calculation is that at the quark level there is a tendency for cancellation between the loop and exchange current contributions and the tendency for cancellation between pionic and vector meson contributions.
Acknowledgement
I thank Professor R. D. McKeown for his hospitality at the W. K. Kellogg Radiation Laboratory of the California Institute of Technology during the completion of this work. Research supported by the Academy of Finland under contract 43982.
Table 1
Flavor factors in the loop integrals, when the parity conserving coupling preceeds the parity violating one. The factors with an asterisk change sign when the ordering of the couplings is reversed.
| $`u\pi ^+du`$ | $`2i^{}`$ |
| --- | --- |
| $`u\pi ^0uu`$ | $`0`$ |
| $`d\pi ^{}ud`$ | $`2i^{}`$ |
| $`d\pi ^0dd`$ | $`0`$ |
| $`u\rho ^+du`$ | $`2(h_{\rho qq}^0)`$ |
| | $`0(h_{\rho qq}^1)`$ |
| | $`1/\sqrt{6}(h_{\rho qq}^2)`$ |
| $`u\rho ^0uu`$ | 1 $`(h_{\rho qq}^0)`$ |
| | $`1(h_{\rho qq}^1)`$ |
| | $`1/\sqrt{6}(h_{\rho qq}^2)`$ |
| $`d\rho ^{}ud`$ | $`2(h_{\rho qq}^0)`$ |
| | $`1/\sqrt{6}(h_{\rho qq}^2)`$ |
| $`d\rho ^0dd`$ | $`1(h_{\rho qq}^0)`$ |
| | $`1(h_{\rho qq}^1)`$ |
| | $`1/\sqrt{6}(h_{\rho qq}^2)`$ |
| $`u\rho ^0uu`$ | $`1(h_{\omega qq}^0)`$ |
| | $`1(h_{\omega qq}^1)`$ |
| $`d\rho ^0uu`$ | $`1(h_{\omega qq}^0)`$ |
| | $`1(h_{\omega qq}^1)`$ |
Table 2
Explicit wave functions for the $`\frac{1}{2}^+`$ and $`\frac{1}{2}^{}`$ non-strange baryon states in the lowest $`S`$\- and $`P`$-shells. The functions $`\phi _{nlm}`$ are harmonic oscillator wave functions. The subscripts $`\pm `$ on the spin-isospin states denote the Yamanouchi symbols (112) and (121) respectively .
| $`p,n`$ | $`\frac{1}{\sqrt{2}}\phi _{000}(\stackrel{}{\rho })\phi _{000}(\stackrel{}{r})\{|\frac{1}{2},t>_+|\frac{1}{2},s>_++|\frac{1}{2},t>_{}|\frac{1}{2}s>_{}\}`$ |
| --- | --- |
| $`N(1535)`$ | $`\frac{1}{2}_{m\sigma }(1,\frac{1}{2},m,\sigma |\frac{1}{2}s)\{\phi _{01m}(\stackrel{}{\rho })\phi _{000}(\stackrel{}{r})`$ |
| | $`[\frac{1}{2},t>_+|\frac{1}{2}\sigma >_+|\frac{1}{2},t>_{}|\frac{1}{2},\sigma >_{}]`$ |
| | $`+\phi _{000}(\stackrel{}{\rho })\phi _{01m}(\stackrel{}{r})[|\frac{1}{2},t>_+|\frac{1}{2},\sigma >_{}+|\frac{1}{2},t>_{}|\frac{1}{2},\sigma >_+]\}`$ |
| $`\mathrm{\Delta }(1620)`$ | $`\frac{1}{\sqrt{2}}_{m\sigma }(1,\frac{1}{2},m,\sigma |\frac{1}{2},s)\{\phi _{01m}(\stackrel{}{\rho })\phi _{000}(\stackrel{}{r})|\frac{3}{2},t>|\frac{1}{2},\sigma >_+`$ |
| | $`+\phi _{000}(\stackrel{}{\rho })\phi _{01m}(\stackrel{}{r})|\frac{3}{2},t>|\frac{1}{2},\sigma >_{}\}`$ |
| $`N(1650)`$ | $`\frac{1}{\sqrt{2}}_{m\sigma }(1,\frac{3}{2},m,\sigma |\frac{1}{2},s)\{\phi _{01m}(\stackrel{}{\rho })\phi _{000}(\stackrel{}{r})|\frac{1}{2},t>_+`$ |
| | $`+\phi _{000}(\stackrel{}{\rho })\phi _{01m}(\stackrel{}{r})|\frac{1}{2},t>_{}\}|\frac{3}{2},\sigma >`$ |
Table 3
Spin-flavor matrix elements of exchange current operators for protons and neutrons with $`s_z=+1/2`$.
| Operator | $`<p,\frac{1}{2}|0|p,\frac{1}{2}>`$ | $`<n,\frac{1}{2}|0|n,\frac{1}{2}>`$ |
| --- | --- | --- |
| $`(\stackrel{}{\tau }^1\stackrel{}{\tau }^2\tau _3^1\tau _3^2)(\sigma _3^1+\sigma _3^2)`$ | 4/3 | 4/3 |
| $`(\stackrel{}{\tau }^1\times \stackrel{}{\tau }^2)_3(\stackrel{}{\sigma }^1\times \stackrel{}{\sigma }^2)_3`$ | -4 | 4 |
| $`\sigma _3^1+\sigma _3^2`$ | 2 | 2 |
| $`\sigma _3^1\tau _3^1+\sigma _3^2\tau _3^2`$ | $`\frac{10}{3}`$ | $`\frac{10}{3}`$ |
| $`\sigma _3^1\tau _3^2+\sigma ^2\tau _3^1`$ | $`\frac{2}{3}`$ | $`\frac{10}{3}`$ |
| $`(\sigma _3^1+\sigma _3^2)\tau ^1\tau ^2`$ | 2 | 2 |
Table 4
Mesonic contributions to the anapole moment of the proton and the neutron.
| | proton | neutron |
| --- | --- | --- |
| $`\pi `$ loops | $`+3.2210^8`$ | $`+1.4510^8`$ |
| $`\pi `$ exchange | $`1.1510^8`$ | $`1.1510^8`$ |
| $`\pi `$ polarization | $`0.1910^8`$ | $`0.0210^8`$ |
| $`\rho `$ loops | $`0.4410^8`$ | $`+1.5210^8`$ |
| $`\rho `$ exchange | $`1.7910^8`$ | $`+0.7310^8`$ |
| $`\rho `$ polarization | $`+0.08810^8`$ | $`0.04710^8`$ |
| $`\omega `$ loops | $`0.09710^8`$ | $`+0.3410^8`$ |
| $`\rho \pi `$ loops | $`0.6710^8`$ | $`0.6510^8`$ |
| $`\omega \pi `$ loops | $`+0.1210^8`$ | $`0.03510^8`$ |
| Total | $`0.9010^8`$ | $`+0.6810^8`$ |
Figure Captions
Figure 1. Pion loop contributions to the anapole form factors of the constituent quarks. The square vertex represents the parity violating pion-quark coupling (2.1a). To the current operator should be added the contributions with the PV and PC hadronic vertices permuted.
Figure 2. Self energy contributions to the anapole moments of the constituent quarks that are required for current conservation. The square vertex represents the parity violating pion-quark coupling (2.1a). To these should be added the corresponding contributions with the PV and PC hadronic vertices permuted.
Figure 3. Calculated pion loop contributions to the anapole form factor of the proton and the neutron.
Figure 4. Pion exchange current contributions to the anapole form factors of the nucleons. The fermion lines represent constituent quarks. The square vertex represents the PV pion-quark coupling (2.1a). The inclusion of the terms with the PV and PC hadronic vertices should be understood.
Figure 5. Pion exchange current and polarization current contributions to the anapole form factor of the proton. The PV pion exchange current contribution is denoted โPI EXCโ and the pion exchange induced polarization current contribution is denoted PI POL. The curves A and B where obtained with the values $`\omega =`$311 MeV and $`\omega =`$1240 MeV respectively for the quark wave function parameter $`\omega `$ (3.5)
Figure 6. Pion exchange current and polarization current contributions to the anapole form factor of the neutron. The PV pion exchange current contribution is denoted โPI EXCโ and the pion exchange induced polarization current contribution is denoted PI POL. The curves A and B where obtained with the values $`\omega =`$311 MeV and $`\omega =`$1240 MeV respectively for the quark wave function parameter $`\omega `$ (3.5)
Figure 7. Schematic representation of the pion polarization current that arises from the PV pion exchange interaction (3.3) induced negative parity admixture in the nucleon wave function. The presence additional terms with the PC and PV pion-quark vertices permuted should be understood.
Figure 8. Combined pion exchange current and polarization current contributions to the anapole form factors of the proton and the neutron.
Figure 9. Vector meson loop contributions to the anapole form factors of constituent quarks. The square vertex represents the parity-violating couplings of vector mesons to constituent quarks ((4.2) and (4.12). These should be combined with the loop amplitudes with the PV and PC hadronic vertices permuted.
Figure 10. $`\rho `$ and $`\omega `$ meson loop contributions to the anapole form factors of the proton and the neutron.
Figure 11. Parity violating $`\rho `$ meson exchange current contributions to the anapole form factors of the nucleon. The square vertex represents the PV $`\rho `$ meson coupling to constituent quarks (4.2). The inclusion of the terms with the PV and PC hadronic vertices should be understood.
Figure 12. $`\rho `$meson exchange current (EXC) and polarization current contributions (POL) to the anapole form factor of the proton. The curves A and B where obtained with the values $`\omega =`$1250 MeV and $`\omega =`$311 MeV respectively for the quark wave function parameter $`\omega `$ (3.5).
Figure 13. $`\rho `$meson exchange current (EXC) and polarization current contributions (POL) to the anapole form factor of the neutron. The curves A and B where obtained with the values $`\omega =`$1250 MeV and $`\omega =`$311 MeV respectively for the quark wave function parameter $`\omega `$ (3.5).
Figure 14. Parity violating $`\rho \pi `$ and $`\omega \pi `$ loop contributions to the anapole form factors of the nucleons. The square vertex represents the parity violating pion-quark coupling (2.1a).
Figure 15. The $`\rho \pi `$ and $`\omega \pi `$ loop contributions to the anapole form factors of the proton and the neutron.
Figure 16. Mixed vector-pseudoscalar exchange current operators.
Figure 17. The net meson loop and exchange current contributions to the anapole form factor of the proton and the neutron. The results calculated without constituent quark form factors are shown separately. |
warning/0003/hep-th0003071.html | ar5iv | text | # Introduction
## Introduction
The presence of a negative cosmological constant is enough to invalidate the classical theorems in which it is proven that at any given time black-hole horizons are always topologically spheres: asymptotically anti-De Sitter ($`aaDS`$) black-hole solutions are known such that the constant-time sections of their event horizons are not topologically spheres . In particular, $`aaDS`$ Schwarzschild black holes with horizons with the topology of Riemann surfaces of arbitrary genus (henceforth called topological black holes) were given in Ref. , the charged generalization in the framework of the Einstein-Maxwell theory with a negative cosmological constant (topological $`aaDS`$ Reissner-Nordstrรถm (RN-$`aDS`$) black holes) was studied in Ref. . The generalization to the rotating case (topological $`aaDS`$ Kerr-Newman (KN-$`aDS`$) black holes) was found and studied in Ref. using the general Petrov type D solution of Plebanski and Demianski (PD solution) Ref. (which contains in different limits all these topological black-hole solutions) and other methods. $`aaDS`$ black holes with exotic horizons with topologies are also known in higher dimensions , in theories with dilaton and Lovelock gravity .
The supersymmetry properties of $`aaDS`$ black holes were first studied by Romans in the context of $`N=2,d=4`$ gauged supergravity for RN-$`aDS`$ black holes with spherical horizons. Later on, Kosteleckรฝ and Perry studied the supersymmetry properties of KN-$`aDS`$ black holes . Recently, Caldarelli and Klemm extended Romansโ results to the case of topological RN-$`aDS`$ black holes and extended and corrected Kosteleckรฝ and Perryโs in the spherical KN-$`aDS`$ case in Ref. .
The supersymmetry properties known are far from being understood. In the recent years we have learned how to interpret many supersymmetric solutions as intersections of โelementaryโ supersymmetric solutions preserving half of the supersymmetries. Each additional object in the intersection breaks an additional half of the remaining supersymmetry <sup>4</sup><sup>4</sup>4Except in Hanany-Witten-like cases in which one can add one more object to an intersection without breaking any further supersymmetry. Needless to say that here we use โobjectโ in a loose and general way that may include gravitational instantons, certain kinds of singularities, etc.. Thus, in $`N=2,d=4`$ ungauged supergravity there is essentially one kind of object which is point-like and that breaks a half of the available supersymmetry and one can either break all the supersymmetry or just one half or nothing at all.
In $`N=2,d=4`$ gauged supergravity, however, Romans discovered solutions that preserve just $`1/4`$ of the supersymmetry, characterized by a magnetic charge inversely proportional to the coupling constant. The simplest of those solutions only has magnetic charge (zero mass and electric charge) equal to the minimal amount of magnetic charge allowed by Diracโs quantization condition. It is really difficult to understand this fact using the paradigm of intersection of elementary objects.
Our goal in this paper is to try to gain some insight into this problem by examining more general cases an calculating, if possible, the amount of supersymmetries preserved by the solutions. Thus, in this letter we first present topological Kerr-Newman-Taub-NUT-$`aDS`$ solutions and cosmological generalizations of the Robinson-Bertotti solution and then study their supersymmetry properties together with those of the general Plebanski-Demianski solution from which all of them can be obtained through different contractions. We will see that, generically, these solutions preserve only $`1/4`$ of the available supersymmetries in presence of angular momentum. Our second main result will be the identification of a sort of electric-magnetic duality symmetry of the supersymmetric Plebanski-Demianski solutions that involves the mass and NUT charge.
This paper is organized as follows: in Section 1 we describe $`N=2,d=4`$ gauged Supergravity. In Section 2 we describe the solutions whose supersymmetry properties we are going to study. In Section 3 we study the integrability conditions of the Killing spinor equation for the topological KN-TN-$`aDS`$ solutions. In Section 3.3 and Section 3.4 we perform the same analysis for RB-$`aDS`$ and the general PD solutions respectively. Section 4 contains our conclusions.
## 1 Cosmological EM Theory and $`N=2,d=4`$ Gauged Supergravity
The $`N=2,d=4`$ supergravity multiplet consists of the Vierbein, a couple of real gravitini and a vector field
$$\{e_\mu {}_{}{}^{a},\psi _\mu =\left(\begin{array}{c}\psi _\mu ^1\\ \psi _\mu ^2\end{array}\right),A_\mu \},$$
(1.1)
respectively. With this multiplet one can construct two different supergravity theories: โpureโ $`N=2,d=4`$ supergravity and โgaugedโ $`N=2,d=4`$ supergravity. The former can be understood as the zero-coupling limit of the latter and the second as the theory one obtains by gauging the $`SO(2)`$ symmetry that rotates the gravitini. The gauged $`N=2,d=4`$ supergravity action for these fields in the 1.5 formalism is
$$\begin{array}{ccc}\hfill S_g& =& d^4xe\{R(e,\omega )+6g^2+2e^1ฯต^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\gamma _\nu (\widehat{๐}_\rho +igA_\rho \sigma ^2)\psi _\sigma ^2\hfill \\ & & \\ & & +๐ฅ_{(m)}{}_{}{}^{\mu \nu }(๐ฅ_{(e)\mu \nu }+๐ฅ_{(m)\mu \nu })\},\hfill \end{array}$$
(1.2)
where $`\widehat{๐}`$ is the $`SO(2,3)`$ gauge covariant derivative
$$\widehat{๐}_\mu =๐_\mu \frac{i}{2}g\gamma _\mu ,$$
(1.3)
$`F`$ is the standard vector field strength, $`\stackrel{~}{F}`$ is the supercovariant field strength and we also define for convenience $``$ by
$$\{\begin{array}{ccc}\hfill F_{\mu \nu }& =& 2_{[\mu }A_{\nu ]},\hfill \\ & & \\ \hfill \stackrel{~}{F}_{\mu \nu }& =& F_{\mu \nu }+๐ฅ_{(e)\mu \nu },\hfill \\ & & \\ \hfill _{\mu \nu }& =& \stackrel{~}{F}_{\mu \nu }+๐ฅ_{(m)\mu \nu },\hfill \end{array}$$
(1.4)
where we have also defined
$$\{\begin{array}{ccc}\hfill ๐ฅ_{(e)\mu \nu }& =& i\overline{\psi }_\mu \sigma ^2\psi _\nu ,\hfill \\ & & \\ \hfill ๐ฅ_{(m)\mu \nu }& =& \frac{1}{2e}ฯต^{\mu \nu \rho \sigma }\overline{\psi }_\rho \gamma _5\sigma ^2\psi _\sigma .\hfill \end{array}$$
(1.5)
We see that the gauge coupling constant $`g`$ is related to the cosmological constant by
$$\mathrm{\Lambda }=3g^2.$$
(1.6)
The equation of motion for $`\omega _\mu ^{ab}`$ implies that it is given by
$$\{\begin{array}{ccc}\hfill \omega _{abc}& =& \mathrm{\Omega }_{abc}+\mathrm{\Omega }_{bca}\mathrm{\Omega }_{cab},\hfill \\ & & \\ \hfill \mathrm{\Omega }_{\mu \nu }^a& =& \mathrm{\Omega }_{\mu \nu }{}_{}{}^{a}(e)+\frac{1}{2}T_{\mu \nu }{}_{}{}^{a},\hfill \\ & & \\ \hfill \mathrm{\Omega }_{abc}(e)& =& e^\mu {}_{a}{}^{}e_{}^{\nu }{}_{b}{}^{}_{[\mu }^{}e_{\nu ]c},\hfill \\ & & \\ \hfill T_{\mu \nu }^a& =& i\overline{\psi }_\mu \gamma ^a\psi _\nu .\hfill \end{array}$$
(1.7)
It is assumed that this equation has been used everywhere (1.5 formalism).
The Maxwell equation and Bianchi identity are
$$\{\begin{array}{ccc}\hfill _\mu (e^{\mu \nu })& =& \frac{ig}{2}ฯต^{\nu \lambda \rho \sigma }\overline{\psi }_\lambda \gamma _5\gamma _\rho \sigma ^2\psi _\sigma ,\hfill \\ & & \\ \hfill _\mu (e{}_{}{}^{}F_{}^{\mu \nu })& =& 0.\hfill \end{array}$$
(1.8)
Observe that the divergences of $`๐ฅ_e`$ and $`๐ฅ_m`$ are two topologically conserved currents that appear as electric-like and magnetic-like sources for the vector field in the Maxwell equation
$$_\mu (eF^{\mu \nu })=+_\mu (e๐ฅ_e^{\nu \mu })+_\mu (e๐ฅ_m^{\nu \mu })+\frac{ig}{2}ฯต^{\nu \lambda \rho \sigma }\overline{\psi }_\lambda \gamma _5\gamma _\rho \sigma ^2\psi _\sigma .$$
(1.9)
They are naturally associated to the electric and magnetic central charges of the $`N=2,d=4`$ supersymmetry algebra. The third term in the r.h.s. of the above equation is associated to the gravitino electric charge and it is, therefore, proportional to the gauge constant. Finally, the Einstein and Rarita-Schwinger equations are
$$\{\begin{array}{ccc}\hfill 0& =& G_a{}_{}{}^{\mu }3g^2e_a{}_{}{}^{\mu }2T(\psi )_a{}_{}{}^{\mu }2\stackrel{~}{T}(A)_a{}_{}{}^{\mu },\hfill \\ & & \\ \hfill 0& =& e^1ฯต^{\mu \nu \rho \sigma }\gamma _5\gamma _\nu \left(\widehat{๐}_\rho +igA_\rho \sigma ^2\right)\psi _\sigma i\left(\stackrel{~}{F}^{\mu \nu }+i{}_{}{}^{}\stackrel{~}{F}_{}^{\mu \nu }\gamma _5\right)\sigma ^2\psi _\nu ,\hfill \end{array}$$
(1.10)
where the equation of motion for $`\omega _\mu ^{ab}`$ has been used and where
$$\{\begin{array}{ccc}\hfill T(\psi )_a^\mu & =& \frac{1}{2e}ฯต^{\mu \nu \rho \sigma }\overline{\psi }_\nu \gamma _5\gamma _a\left(\widehat{๐}_\rho +igA_\rho \sigma ^2\right)\psi _\sigma \hfill \\ & & \\ & & \frac{ig}{4e}ฯต^{\mu \nu \rho \sigma }\overline{\psi }_\nu \gamma _5\gamma _{\rho a}\psi _\sigma ,\hfill \\ & & \\ \hfill \stackrel{~}{T}(A)_a^\mu & =& \stackrel{~}{F}_a{}_{}{}^{\rho }\stackrel{~}{F}_{}^{\mu }{}_{\rho }{}^{}\frac{1}{4}e_a{}_{}{}^{\mu }\stackrel{~}{F}_{}^{2}.\hfill \end{array}$$
(1.11)
Apart from invariance under general coordinate and local Lorentz transformations the theory is invariant under $`U(1)`$ gauge transformations
$$\{\begin{array}{ccc}\hfill A_\mu ^{}& =& A_\mu +_\mu \chi ,\hfill \\ & & \\ \hfill \psi _\mu ^{}& =& e^{ig\chi \sigma ^2}\psi _\mu ,\hfill \end{array}$$
(1.12)
and local $`N=2`$ supersymmetry transformations
$$\{\begin{array}{ccc}\hfill \delta _ฯตe_\mu ^a& =& i\overline{ฯต}\gamma ^a\psi _\mu ,\hfill \\ & & \\ \hfill \delta _ฯตA_\mu & =& i\overline{ฯต}\sigma ^2\psi _\mu ,\hfill \\ & & \\ \hfill \delta _ฯต\psi _\mu & =& \stackrel{~}{\widehat{๐}}_\mu ฯต,\hfill \end{array}$$
(1.13)
where the $`\stackrel{~}{\widehat{๐}}_\mu `$ is the supercovariant derivative defined by
$$\stackrel{~}{\widehat{๐}}_\mu =\widehat{๐}_\mu +igA_\mu \sigma ^2+\frac{1}{4}\overline{)}\stackrel{~}{F}\gamma _\mu \sigma ^2.$$
(1.14)
In the ungauged case, the theory enjoys chiral-dual invariance which interchanges the Maxwell and Bianchi equations and the topologically conserved electric and magnetic charges (and, therefore, the associated central charges). In the gauged theory, the gauge coupling breaks this invariance.
We are going to work with purely bosonic solutions of this theory. They obey the bosonic equations of motion
$$\{\begin{array}{ccc}\hfill _\mu F^{\mu \nu }& =& 0,\hfill \\ & & \\ \hfill _\mu {}_{}{}^{}F_{}^{\mu \nu }& =& 0,\hfill \\ & & \\ \hfill R_{\mu \nu }& =& 2T_{\mu \nu }(A)3g^2g_{\mu \nu },\hfill \end{array}$$
(1.15)
where $`T_{\mu \nu }(A)`$ is just the standard energy-momentum tensor for an Abelian gauge field:
$$T_{\mu \nu }(A)=F_\mu {}_{}{}^{\rho }F_{\rho \nu }^{}\frac{1}{4}g_{\mu \nu }F^2.$$
(1.16)
These equations of motion are duality-invariant. However, the gravitino supersymmetry rule (even with fermionic fields set to zero) is not duality-invariant and the supersymmetry properties of duality-related bosonic solutions are not, in general, the same.
## 2 Topological RN-TN-$`aDS`$, KN-TN-$`aDS`$ and RB-$`aDS`$ and PD Solutions
In this section we display and describe the solutions whose supersymmetry properties will later be studied. For simplicity we start with the unrotating RN-TN-$`aDS`$ although they are included in the general KN-TN-$`aDS`$ case.
### 2.1 Topological RN-TN-$`aDS`$ Solutions
These solutions generalize, by including NUT charge $`N`$, the topological RN-$`aDS`$ black hole solutions found in Ref. . There are three cases labeled by the parameter $`\mathrm{}`$ whose value is essentially the sign of one minus the genus of the horizon and therefore takes the values $`1,0,1`$ for the sphere (genus zero), the torus (genus 1) and higher genus Riemann surfaces, respectively. In the three cases the metric can be written in this form
$$\{\begin{array}{ccc}\hfill ds^2& =& \frac{\lambda }{R^2}\left(dt+\omega _{\mathrm{}}d\phi \right)^2\frac{R^2}{\lambda }dr^2R^2d\mathrm{\Omega }_{\mathrm{}}^2,\hfill \\ & & \\ \hfill \lambda & =& \left[g^2R^4+(\mathrm{}+4g^2N^2)(r^2N^2)2Mr+|Z|^2\right],\hfill \\ & & \\ \hfill R^2& =& r^2+N^2,\hfill \end{array}$$
(2.1)
where $`d\mathrm{\Omega }_{\mathrm{}}^2`$ is the metric of the unit sphere, the plane and the upper half plane respectively
$$d\mathrm{\Omega }_{\mathrm{}}^2=\{\begin{array}{cccc}d\theta ^2+\mathrm{sin}^2\theta d\phi ^2,\hfill & \hfill \mathrm{}& =& +1,\hfill \\ & & & \\ d\theta ^2+d\phi ^2,\hfill & \hfill \mathrm{}& =& 0,\hfill \\ & & & \\ d\theta ^2+\mathrm{sinh}^2\theta d\phi ^2,\hfill & \hfill \mathrm{}& =& 1,\hfill \end{array}$$
(2.2)
$`\omega _{\mathrm{}}`$ is the function
$$\omega _{\mathrm{}}=\{\begin{array}{cccc}2N\mathrm{cos}\theta ,\hfill & \hfill \mathrm{}& =& +1,\hfill \\ & & & \\ 2N\theta ,\hfill & \hfill \mathrm{}& =& 0,\hfill \\ & & & \\ 2N\mathrm{cosh}\theta ,\hfill & \hfill \mathrm{}& =& 1,\hfill \end{array}$$
(2.3)
and the vector potential is given by
$$\begin{array}{ccc}\hfill A_t& =& \left(QrNP\right)/R^2,\hfill \\ & & \\ \hfill A_\phi & =& \{\begin{array}{ccccc}\mathrm{cos}\theta \left[P(r^2N^2)+2NQr\right]/R^2,\hfill & \mathrm{for}& \hfill \mathrm{}& =& +1,\hfill \\ & & & & \\ \theta \left[P(r^2N^2)+2NQr\right]/R^2,\hfill & \mathrm{for}& \hfill \mathrm{}& =& 0,\hfill \\ & & & & \\ \mathrm{cosh}\theta \left[P(r^2N^2)+2NQr\right]/R^2,\hfill & \mathrm{for}& \hfill \mathrm{}& =& 1.\hfill \end{array}\hfill \end{array}$$
(2.4)
It is understood that one has to take the equation of the last two spacetimes by a discrete group in order to get a torus or a Riemann surface of arbitrary genus.
These solutions are valid in the $`g=0`$ case. In that limit (with $`N=0`$) so we can speak of black holes, only the $`\mathrm{}=+1`$ ones can have a regular event horizon, in agreement with . With $`g0`$ (still with $`N=0`$) and we recover the solutions of Ref. in which $`M`$ is the mass, $`Q`$ the electric charge, $`P`$ the magnetic charge and $`Z=Q+iP`$ some of which are black holes with regular horizons of different topologies.
For $`g=0,N0`$ we recover the standard RN-TN solutions in which those parameters are still the physical parameters<sup>5</sup><sup>5</sup>5A definition of the mass of Taub-NUT spaces cannot be given in the standard form because these solutions do not go asymptotically to any other vacuum solution. The same happens in the 5-dimensional KK monopole solution, studied in Refs. . However, as different from the KK monopole, the TN solution is not ultrastatic and the tricks used in those references to define and calculate the mass of the KK monopole do not seem to apply to this case. A definition inspired in the AdS/CFT correspondence, has, however, been recently given in Refs. . and $`N`$ is the NUT charge. When the product $`gN0`$ it is no longer clear that $`M,Q,P`$ are the true mass, electric and magnetic charges that appear in the superalgebra. This is similar to what happens in the rotating case in which the true charges are combinations of the parameters $`M,P,Q`$ appearing in the solution with the product $`ga`$.
It is useful to have a general form of the solutions valid for the three cases $`\mathrm{}=1,0,1`$. To have such a general expression we define the coordinate $`u`$
$$u\{\begin{array}{cccc}\mathrm{cos}\theta ,\hfill & \hfill \mathrm{}& =& +1,\hfill \\ & & & \\ \theta ,\hfill & \hfill \mathrm{}& =& 0,\hfill \\ & & & \\ \mathrm{cosh}\theta ,\hfill & \hfill \mathrm{}& =& 1,\hfill \end{array}$$
(2.5)
and then
$$\{\begin{array}{ccc}\hfill ds^2& =& \frac{\lambda }{R^2}\left(dt2Nud\phi \right)^2\frac{R^2}{\lambda }dr^2\frac{R^2}{S(u)}du^2R^2S(u)d\phi ^2,\hfill \\ & & \\ \hfill A_t& =& \left(QrNP\right)/R^2,\hfill \\ & & \\ \hfill A_\phi & =& u\left[P\left(r^2N^2\right)+2NrQ\right]/R^2,\hfill \\ & & \\ \hfill S(u)& =& \mathrm{}(1u^2)+1\mathrm{}^2,\hfill \end{array}$$
(2.6)
where $`\lambda `$ and $`R`$ are as above.
### 2.2 Topological KN-TN-$`aDS`$ Solutions
These solutions generalize the topological KN-$`aDS`$ solutions given in Ref. to the non-zero NUT charge case. In the $`t,r,u,\phi `$ coordinate system (which is Boyer-Lindquist-type) they can be written as follows:
$$\{\begin{array}{ccc}\hfill ds^2& =& \frac{\lambda }{R^2(r,u)}\left\{dt\left[2Nua\left(\mathrm{}^2u^2\right)\right]d\phi \right\}^2\frac{R^2(r,u)}{\lambda }dr^2\hfill \\ & & \\ & & \frac{R^2(r,u)}{๐ฎ(u)}du^2\frac{๐ฎ(u)}{R^2(r,u)}\left[\left(r^2+N^2+\mathrm{}^2a^2\right)d\phi +adt\right]^2.\hfill \\ & & \\ \hfill A_t& =& \left[QrP(N+au)\right]/R^2(r,u),\hfill \\ & & \\ \hfill A_\phi & =& \frac{1}{a}\sqrt{r^2+N^2+\mathrm{}^2a^2}\left[QrP(N+au)\right]/R^2(r,u),\hfill \\ & & \\ \hfill \lambda & =& g^2r^4+\left(\mathrm{}+\mathrm{}^2a^2g^2+6g^2N^2\right)r^22Mr+|Z|^2\hfill \\ & & \\ & & N^2\left(\mathrm{}3\mathrm{}^2a^2g^2+3g^2N^2\right)+a^2\left(1+\mathrm{}\mathrm{}^2\right),\hfill \\ & & \\ \hfill ๐ฎ(u)& =& S(u)+\left(a^2g^2u^2+\mathrm{\hspace{0.17em}4}ag^2Nu\right)\left(u^2\mathrm{}^2\right),\hfill \\ & & \\ \hfill R^2(r,u)& =& r^2+\left(N+au\right)^2,\hfill \end{array}$$
(2.7)
with $`S(u)`$ as above.
In Appendix A it is explained how this metric can be obtained from the general solution of Plebanski and Demianski . The above form of the potentials is valid for $`a0`$. The $`a0`$ limit of the field strength is perfectly well defined.
### 2.3 Topological RB Solutions
In ungauged $`N=2,d=4`$ Supergravity, the extremal RN black hole can be seen as a soliton interpolating between two supersymmetric vacua: Minkowski spacetime at infinity and RB in the near-horizon limit. The RB spacetime is the product $`aDS_2\times S^2`$ where both factors are maximally symmetric spaces with opposite curvatures that cancel each other. The same thing occurs with other $`p`$-branes in higher dimensions where the role of the RB spacetime is played by $`aDS_{p+2}\times S^{8p}`$. Here we present a generalization of the RB spacetime to the case of gauged $`N=2,d=4`$ Supergravity (cosmological E-M theory) whose supersymmetry properties we will study later. They are the product of $`aDS_2`$ with a sphere $`S^2`$, a torus $`T^2`$ or a higher-genus Riemann surface $`\mathrm{\Sigma }_g`$ in which now the curvature of the $`aDS_2`$ spacetime is not completely canceled by the other factor space but they add up to the 4-dimensional cosmological constant
$$\{\begin{array}{ccc}\hfill ds^2& =& K^2r^2dt^2\frac{1}{K^2r^2}dr^2L^2S(u)^1du^2L^2S(u)d\phi ^2,\hfill \\ & & \\ \hfill F_{01}& =& \alpha \hfill \\ & & \\ \hfill F_{23}& =& \beta ,\hfill \end{array}$$
(2.8)
where the constants $`K,L,\alpha ,\beta `$ satisfy
$$\begin{array}{ccc}\hfill g^2& =& \frac{1}{6}\left\{K^2\mathrm{}L^2\right\},\hfill \\ & & \\ \hfill \alpha ^2+\beta ^2& =& \frac{1}{2}\left(K^2+\mathrm{}L^2\right).\hfill \end{array}$$
(2.9)
The field strength is covariantly constant and in this coordinate system has constant components which correspond to the vector potential components
$$\{\begin{array}{ccc}\hfill A_t& =& \alpha r,\hfill \\ & & \\ \hfill A_\phi & =& \beta /L^2u.\hfill \end{array}$$
(2.10)
The $`\mathrm{}=1,K^2=2L^2`$ solution, which has special supersymmetry properties has been also given in .
### 2.4 PD Solutions
Plebanski and Demianski found most general Petrov type D solution of the cosmological E-M theory. This general solution contains as limiting cases all the known solutions, and, in particular the topological KN-TN-$`aDS`$ solutions presented above (which in their turn, also contain the RN-TN-$`aDS`$ solutions presented at the beginning). This is shown in Appendix A.
The PD solution depends on the constants<sup>6</sup><sup>6</sup>6These constants are different from the constants $`M,N,Q,P`$ that appear in the previous solutions. $`๐ฌ,๐ญ,๐ฐ,๐ฏ,๐ค,\alpha `$ and, of course, $`g`$, and, in Boyer coordinates $`\tau ,\sigma ,p,q`$, reads
$$\{\begin{array}{ccc}\hfill ds^2& =& \frac{๐ฌ(q)}{p^2+q^2}\left(d\tau p^2d\sigma \right)^2\frac{p^2+q^2}{๐ฌ(q)}dq^2\frac{p^2+q^2}{๐ซ(p)}dp^2\frac{๐ซ(p)}{p^2+q^2}\left(d\tau +q^2d\sigma \right)^2,\hfill \\ & & \\ \hfill F_{01}& =& (q^2+p^2)^2\left[๐ฐ(q^2p^2)2๐ฏpq\right],\hfill \\ & & \\ \hfill F_{23}& =& (q^2+p^2)^2\left[๐ฏ(q^2p^2)+2๐ฐpq\right],\hfill \\ & & \\ \hfill ๐ฌ(q)& =& g^2q^4+๐คq^22๐ฌq+๐ฐ^2+\alpha ,\hfill \\ & & \\ \hfill ๐ซ(p)& =& g^2p^4๐คp^2+2๐ญp๐ฏ^2+\alpha .\hfill \end{array}$$
(2.11)
This class of solutions has a scaling invariance given by
$$\begin{array}{ccccccccc}\hfill q& & \kappa q,\hfill & \hfill ๐ฌ& & \kappa ^3๐ฌ,\hfill & \hfill ๐ญ& & \kappa ^3๐ญ,\hfill \\ & & & & & & & & \\ \hfill p& & \kappa p,\hfill & \hfill ๐ฐ& & \kappa ^2๐ฐ,\hfill & \hfill ๐ฏ& & \kappa ^2๐ฏ,\hfill \\ & & & & & & & & \\ \hfill \tau & & \kappa ^1\tau ,\hfill & \hfill ๐ค& & \kappa ^2๐ค,\hfill & \hfill \alpha & & \kappa ^4\alpha ,\hfill \\ & & & & & & & & \\ \hfill \sigma & & \kappa ^3\sigma ,\hfill & & & & & & \end{array}$$
(2.12)
which can be used to bring one of the charges to a given value. It is clear that this scaling freedom remains if one of the charges happens to be nil.
The curvature is determined by $`๐ฌ`$, $`๐ญ`$, $`๐ฐ`$ and $`๐ฏ`$ and one can see that when they are zero, the Weyl tensor vanishes. This then means that in that case, the solution is locally $`aDS_4`$.
## 3 Supersymmetry and Integrability Conditions
The bosonic part of the supercovariant derivative for gauged $`N=2`$ supergravity is
$$\stackrel{~}{\widehat{๐}}_\mu =\widehat{}_\mu +gA_\mu i\sigma ^2\frac{i}{4}\overline{)}F\gamma _\mu i\sigma ^2,$$
(3.1)
where $`\widehat{}_\mu `$ is the $`SO(2,3)`$ gauge-covariant derivative. The Killing spinor equation is
$$\stackrel{~}{\widehat{D}}_\mu ฯต=0,$$
(3.2)
and a necessary condition for it to have solutions is the integrability condition
$$[\stackrel{~}{\widehat{๐}}_\mu ,\stackrel{~}{\widehat{๐}}_\nu ]ฯต=0.$$
(3.3)
One finds
$$\begin{array}{ccc}\hfill [\stackrel{~}{\widehat{๐}}_\mu ,\stackrel{~}{\widehat{๐}}_\nu ]ฯต& =& \frac{1}{4}\{C_{\mu \nu }{}_{}{}^{ab}\gamma _{ab}^{}+2i\overline{)}(F_{\mu \nu }+i{}_{}{}^{}F_{\mu \nu }^{}\gamma _5)i\sigma ^2\hfill \\ & & \\ & & +\frac{g}{2}F_{ab}(3\gamma ^{ab}\gamma _{\mu \nu }+\gamma _{\mu \nu }\gamma ^{ab})i\sigma ^2\}ฯต=0.\hfill \end{array}$$
(3.4)
We study first the non-rotating case RN-TN-$`aDS`$ case.
### 3.1 Supersymmetry of Topological RN-TN-$`aDS`$ Solutions
Introducing the Vierbein 1-forms
$$\begin{array}{cccccc}\hfill e^0& =& \lambda ^{1/2}/R\left(dt+\omega _{\mathrm{}}d\phi \right),\hfill & \hfill e^1& =& \lambda ^{1/2}Rdr,\hfill \\ & & & & & \\ \hfill e^2& =& Rd\theta ,\hfill & \hfill e^3& =& R\mathrm{\Omega }_{\mathrm{}}d\phi ,\hfill \end{array}$$
(3.5)
we find
$$\begin{array}{ccc}\hfill F_{01}& =& \left(Q(r^2N^2)2NPr\right)/R^4,\hfill \\ & & \\ \hfill F_{23}& =& \left(P(r^2N^2)+2NQr\right)/R^4,\hfill \end{array}$$
(3.6)
$$\begin{array}{ccc}\hfill _1F_{01}& =& 2\lambda ^{1/2}/R^7\left[Q(r^33rN^2)P(3r^2NN^3)\right],\hfill \\ & & \\ \hfill _1{}_{}{}^{}F_{01}^{}& =& 2\lambda ^{1/2}/R^7\left[P(r^33rN^2)+Q(3r^2NN^3)\right],\hfill \end{array}$$
(3.7)
(the remaining components of $`_aF_{bc}`$ can be found using the Bianchi identities or the Maxwell equations, which are satisfied) and
$$\begin{array}{ccc}\hfill \frac{1}{2}C_{01}^{01}& =& C_{02}{}_{}{}^{02}=C_{03}{}_{}{}^{03}=C_{12}{}_{}{}^{12}=C_{13}{}_{}{}^{13}=\frac{1}{2}C_{23}{}_{}{}^{23}=C_1,\hfill \\ & & \\ \hfill C_{02}^{13}& =& C_{03}{}_{}{}^{12}=C_{12}{}_{}{}^{03}=C_{13}{}_{}{}^{02}=\frac{1}{2}C_{23}{}_{}{}^{01}=C_2,\hfill \\ & & \\ \hfill C_1& =& [Mr^3(3N^2(\mathrm{}4g^2N^2)+|Z|^2)r^2\hfill \\ & & \\ & & 3N^2Mr+N^2(N^2(\mathrm{}4g^2N^2)+|Z|^2)]/R^6,\hfill \\ & & \\ \hfill C_2& =& N[(\mathrm{}4g^2N^2)r^3+3Mr^2\hfill \\ & & \\ & & (3N^2(\mathrm{}4g^2N^2)+2|Z|^2)rMN^2]/R^6,\hfill \end{array}$$
(3.8)
Plugging all this into the integrability conditions we get the following conditions on the parameters:
$`0`$ $`=`$ $`g\left[MP+QN(\mathrm{}+4g^2N^2)\right],`$ (3.9)
$`0`$ $`=`$ $`_+_{},`$ (3.10)
where we have defined
$$_\pm (MgNQ)^2+N^2(\mathrm{}\pm gP+4g^2N^2)^2(\mathrm{}\pm 2gP+5g^2N^2)|Z|^2.$$
(3.12)
The first condition plays the role of a constraint which is automatically satisfied in the well-known $`g=0`$ case, while the second implies $`_\pm =0`$ which should be the (saturated) Bogomolโnyi bound of gauged $`N=2,d=4`$ supergravity and actually it reduces to the well-known Bogomolโnyi bound of ungauged $`N=2,d=4`$ supergravity in asymptotically flat spaces ($`\mathrm{}=+1`$) generalized so as to include NUT charge (see Refs. ) $`M^2+N^2=Q^2+P^2`$. For $`g=0`$ and arbitrary $`\mathrm{}`$ we get
$$M^2+\mathrm{}^2N^2=\mathrm{}(Q^2+P^2).$$
(3.13)
A detailed analysis of the different cases in which the constrains is satisfied and the Bogomolโnyi bound is saturated gives as a result the four cases represented in Table 1.
The first corresponds evidently to $`aDS_4`$ itself in standard spherical coordinates, which is maximally supersymmetric and preserves all supersymmetries. The second case can be shown to describe, at least locally, $`aDS_4`$ as well (the Weyl tensor vanishes and the space is maximally symmetric). There are, thus, two different values of the parameter $`N`$ that correspond to the same spacetime.
In the third and fourth cases we have taken for the sake of convenience $`Q`$ and $`N`$ as independent parameters. The third case is a generalization to $`gN0`$ of the standard $`M=|Q|`$ case of ungauged $`N=2,d=4`$ supergravity where $`Q`$ is arbitrary which preserves $`1/2`$ of the supersymmetries. Here a non-vanishing magnetic charge proportional to $`N`$ is induced. As a matter of fact, it admits the limits $`g0`$ and/or $`N0`$ with the same amount of supersymmetry preserved.
There are two particularly interesting limits: the often neglected $`g=0,\mathrm{}=0`$ case which (setting $`N=0`$ for simplicity and rescaling the coordinates $`\theta ,\phi `$ which do not represent angles anymore) corresponds to the solution
$$\{\begin{array}{ccc}\hfill ds^2& =& \frac{Q^2}{r^2}dt^2\frac{r^2}{Q^2}\left(dr^2+d\theta ^2+d\phi ^2\right),\hfill \\ & & \\ \hfill A_t& =& \frac{Q}{r}.\hfill \end{array}$$
(3.14)
This solution belongs to the Papapetrou-Majumdar class
$$\{\begin{array}{ccc}\hfill ds^2& =& H^2dt^2H^2d\stackrel{}{x}^{\mathrm{\hspace{0.17em}2}},\hfill \\ & & \\ \hfill A_t& =& \pm H^1,\hfill \\ & & \\ \hfill _{\underset{ยฏ}{i}}_{\underset{ยฏ}{i}}H& =& 0,\hfill \end{array}$$
(3.15)
where the harmonic function $`H`$ has been chosen to depend on only one coordinate $`H=|Q|x`$ and not on $`y,z`$.
The second interesting limit $`Q0`$ also gives a supersymmetric configuration that preserves $`1/2`$ of the supersymmetries with only magnetic and NUT charge and zero mass.
The fourth case in Table 1 preserves only $`1/4`$ of the supersymmetries and only exists for $`g0`$. It is a generalization to $`N0`$ of Romansโ global monopole solution . We see that the presence of both NUT and electric charge implies that the mass parameter has to be finite. On the other hand, it admits the limits $`Q0`$ and/or $`N0`$ with the same amount of supersymmetry preserved.
### 3.2 Supersymmetry of KN-TN-$`aDS`$ Solutions
We choose the Vierbein 1-forms
$$\begin{array}{ccc}\hfill e^0& =& \frac{\lambda ^{1/2}}{R(r,u)}\left[dt\left(2Nua(\mathrm{}^2u^2)\right)d\phi \right],\hfill \\ & & \\ \hfill e^1& =& \frac{R(r,u)}{\lambda ^{1/2}}dr,\hfill \\ & & \\ \hfill e^2& =& \frac{R(r,u)}{๐ฎ^{1/2}(u)}du,\hfill \\ & & \\ \hfill e^3& =& \frac{๐ฎ^{1/2}(u)}{R(r,u)}\left[(r^2+N^2+\mathrm{}^2a^2)d\phi +adt\right],\hfill \end{array}$$
(3.16)
on which the field strength components read
$$\begin{array}{ccc}\hfill F_{01}& =& R(r,u)^4\left[Q(r^2(N+au)^2)\mathrm{\hspace{0.17em}2}Pr(N+au)\right],\hfill \\ & & \\ \hfill F_{23}& =& R(r,u)^4\left\{P(r^2(N+au)^2)+\mathrm{\hspace{0.17em}2}Qr(N+au)\right\},\hfill \end{array}$$
(3.17)
We only need to calculate
$$\begin{array}{ccc}\hfill C_{0101}& =& 2R^6\left[M(r^33rX^2)+N(\mathrm{}\mathrm{}^2a^2g^2+4g^2N^2)(3r^2XX^3)Z^2(r^2X^2)\right],\hfill \\ & & \\ \hfill C_{0123}& =& 2R^6\left[M(3r^2XX^3)+N(\mathrm{}\mathrm{}^2a^2g^2+4g^2N^2)(3rX^2r^3)2Z^2rX\right],\hfill \\ & & \\ \hfill _1F_{01}& =& 2R^7\lambda ^{1/2}\left[Q(r^33rX^2)P(3r^2XX^3)\right],\hfill \\ & & \\ \hfill _2F_{01}& =& 2aR^7๐ฎ^{1/2}\left\{P(r^33rX^2)+Q(3r^2XX^3)\right\},\hfill \end{array}$$
(3.18)
where we used the abbreviation $`X=N+au`$. As in the RN-TN-$`aDS`$ case the other components of the integrability condition turn out to be proportional to the $`01`$ component. From this one obtains the constraint and generalization of the Bogomolโnyi bound
$$\begin{array}{ccc}\hfill 0& =& g\left[MP+NQ(\mathrm{}\mathrm{}^2a^2g^2+4g^2N^2)\right],\hfill \\ & & \\ \hfill 0& =& _+_{},\hfill \end{array}$$
(3.19)
where now
$$\begin{array}{ccc}\hfill _\pm & & M^2+N^2(\mathrm{}\mathrm{}^2a^2g^2+4g^2N^2)[(\mathrm{}+\mathrm{}^2a^2g^2+6g^2N^2)\hfill \\ & & \\ & & \pm 2g\sqrt{a^2(1+\mathrm{}\mathrm{}^2)N^2\left(\mathrm{}3\mathrm{}^2a^2g^2+3g^2N^2\right)}]Z^2,\hfill \end{array}$$
(3.20)
The fact that the bound factorizes into the product $`_+_{}`$ is difficult to see directly from the calculation but easy to deduce from the results we will find in the general PD case. It can be checked that the (saturated) bound obtained is exactly the same, when $`N=0`$, as the one given by Caldarelli and Klemm in Ref. .
We can now try to analyze different solutions to these two equations. This is a very complex problem and it would only make sense to explain in detail a classification of the solutions if the different classes had different amounts of unbroken supersymmetry. However, in all the cases that we have been able to analyze we have not found any single supersymmetric solution with $`a0`$ preserving $`1/2`$ of the supersymmetries. In fact, adding angular momentum to the RN-TN-$`aDS`$ solutions that do preserve $`1/2`$ of the supersymmetries always seems to break an another half leaving only $`1/4`$ unbroken.
For instance, the solution with $`M=Q=0,P=\pm (2g)^1(\mathrm{}\mathrm{}^2a^2g^2+4g^2N^2),\mathrm{}=\pm 1`$ preserves $`1/2`$ with $`a=0`$ and only $`1/4`$ with $`a0`$. The same effect takes place in all the instances studied.
### 3.3 Supersymmetry of Topological RB Solutions
To check supersymmetry of the topological RB solutions we only need
$$C_{0101}=\frac{1}{3}\left\{K^2\mathrm{}L^2\right\}=2g^2,$$
(3.21)
since the vector field strengths is covariantly constant.
The integrability condition then reads
$$g\left[g๐\alpha \gamma ^{01}i\sigma ^2+\beta \gamma ^{23}i\sigma ^2\right]ฯต=\mathrm{\hspace{0.33em}0}.$$
(3.22)
Obviously, for $`g=0`$ one finds Robinson-Bertotti which will not break any supersymmetry. When $`g0`$ however, one finds, just by taking the determinant of the above equation, that one has to satisfy
$$(g\pm \beta )^2+\alpha ^2=\mathrm{\hspace{0.17em}0}\{\begin{array}{ccc}\alpha \hfill & =& 0\hfill \\ \beta \hfill & =& \pm g\hfill \end{array}$$
(3.23)
which then break half of the available supersymmetry. Plugging the above equations into Eq. (2.9), one finds that
$$\mathrm{}=1,K^2=\mathrm{\hspace{0.33em}2}L^2,$$
(3.24)
which means that $`K^2=4g^2`$ and $`L^2=2g^2`$. This is the solution found in Ref. .
We could have found this solution also as the near-horizon limit of the $`\mathrm{}`$ generalization of Romansโ global monopole . In that case we have $`P=\frac{\pm \mathrm{}}{2g}`$ and with $`\mathrm{}=1`$ and all other charges vanishing we find that there is a horizon at $`2g^2r^2=1`$. At this radius the solution can be approximated by
$$\begin{array}{ccc}\hfill ds^2& =& 4g^2r^2dt^2\frac{1}{4g^2r^2}dr^2\frac{1}{2g^2}\left(d\theta ^2+\mathrm{sinh}^2(\theta )d\varphi ^2\right),\hfill \\ & & \\ \hfill F_{23}& =& \frac{\pm 1}{2g}\left(\frac{1}{2g^2}\right)^1=g.\hfill \end{array}$$
(3.25)
which is just the supersymmetric RB-like solution discussed above. We then see that we have supersymmetry enhancement at the horizon from $`1/4`$ to $`1/2`$. Observe that he presence of electric charge would have meant the complete annihilation of supersymmetry at the horizon.
### 3.4 Supersymmetry of the PD General Solution
As in the foregoing cases, one finds that all the components of the integrability condition are equivalent, so we will only write down the components of the Weyl tensor and the covariant derivative of the vector field strength to calculate the integrability condition in the $`01`$ direction.
$$\begin{array}{ccc}\hfill C_{1010}& =& \frac{2(p+q)^3}{(1+p^2q^2)^3}\left[๐ฌ(13p^2q^2)+๐ญ(3pqp^3q^3)+๐น^2(pq)(1p^2q^2)\right],\hfill \\ & & \\ \hfill C_{1023}& =& \frac{2(p+q)^3}{(1+p^2q^2)^3}\left[๐ฌ(3pq3p^3q^3)๐ญ(13p^2q^2)+๐น^22pq(pq)\right],\hfill \\ & & \\ \hfill _1F_{01}& =& \frac{2(p+q)^2๐ฌ^{1/2}}{(1+p^2q^2)^{7/2}}\left[๐ฐ(13p^3q+p^5q^33p^2q^2)+๐ฏ(3pqp^3q^3+p^23p^4q^2)\right],\hfill \\ & & \\ \hfill _2F_{01}& =& \frac{2(p+q)^2๐ซ^{1/2}}{(1+p^2q^2)^{7/2}}\left[๐ฏ(13p^3q+p^5q^33p^2q^2)๐ฐ(3pqp^3q^3+p^23p^4q^2)\right],\hfill \end{array}$$
(3.26)
Plugging these expressions into the integrability condition and calculating the determinant, one finds that the following conditions need to be satisfied in order for the solution to be supersymmetric
$$\begin{array}{ccc}\hfill 0& =& g\left[\mathrm{๐ฌ๐ฏ}+\mathrm{๐ญ๐ฐ}\right],\hfill \\ & & \\ \hfill 0& =& _+_{},\hfill \end{array}$$
(3.27)
where, now
$$_\pm ๐ถ^2(๐ค\pm 2g\alpha ^{1/2})๐น^2,$$
(3.28)
and we have defined $`๐ถ^2=๐ฌ^2+๐ญ^2`$ and $`๐น^2=๐ฐ^2+๐ฏ^2`$. One can check that these conditions are invariant under the scalings in Eq. (2.12) and they give the integrability equations of the RN-TN-$`aDS`$ and KN-TN-$`aDS`$ cases after the redefinitions (A.2).
Again we find a constraint on the charges and a generalization of the (saturated) Bogomolโnyi bound $`_\pm =0`$. The advantage of the parameterization of the PD solution is, first of all, that the second integrability condition factorizes completely and that $`_\pm `$ is extremely simple and is almost identical to the standard bound for asymptotically flat, ungauged, $`N=2,d=4`$ supergravity solutions, being electric-magnetic duality-invariant and invariant under gravito-electric-magnetic duality that rotates $`๐ฌ`$ into $`๐ญ`$ and vice-versa. These duality invariances are broken by the constraint $`g\left[\mathrm{๐ฌ๐ฏ}+\mathrm{๐ญ๐ฐ}\right]=0`$ which is, nevertheless invariant under simultaneous rotations with the same angle
$$\{\begin{array}{ccc}\hfill ๐ฌ^{}& =& \mathrm{cos}\theta ๐ฌ\mathrm{sin}\theta ๐ญ,\hfill \\ & & \\ \hfill ๐ญ^{}& =& \mathrm{sin}\theta ๐ฌ+\mathrm{cos}\theta ๐ญ,\hfill \end{array}\{\begin{array}{ccc}\hfill ๐ฐ^{}& =& \mathrm{cos}\theta ๐ฐ+\mathrm{sin}\theta ๐ฏ,\hfill \\ & & \\ \hfill ๐ฏ^{}& =& \mathrm{sin}\theta ๐ฐ+\mathrm{cos}\theta ๐ฏ.\hfill \end{array}$$
(3.29)
Actually, assuming that $`g0`$ one can eliminate completely the constraint, getting a pair of equations
$$\{\begin{array}{ccc}\hfill ๐ฌ^2& =& (๐ค\pm 2g\alpha ^{1/2})๐ฐ^2,\hfill \\ & & \\ \hfill ๐ญ^2& =& (๐ค\pm 2g\alpha ^{1/2})๐ฏ^2,\hfill \end{array}$$
(3.30)
which hold even if some of these charges (but not $`g`$) vanish. These equations rotate into each other under the above duality transformations.
The rotation parameter is always bounded above:
$$\alpha ^{1/2}\pm ๐ค/2g.$$
(3.31)
When this bound is saturated, then both $`๐ฌ=0`$ and $`๐ญ=0`$, while $`๐ฐ`$ and $`๐ฏ`$ remain arbitrary. This is always the case when $`๐ค=0`$.
Finally, the only supersymmetric solution with $`๐น=0`$ is $`aDS_4`$.
A calculation of the rank of the integrability condition shows that all these configurations will generically break three-fourths of the available supersymmetries. This was to be expected from our results in the KN-TN-$`aDS`$ case. On the other hand we have not been able to find any combination preserving up to $`1/2`$ of the available supersymmetry which is not the RN-TN-$`aDS`$ solution.
## 4 Conclusions
In this letter we have presented new solutions which generalize the already known topological black holes and the standard Robinson-Bertotti solution. We have explored their supersymmetry properties finding that generically they preserve only $`1/4`$ of the supersymmetry. The only solutions that preserve $`1/2`$ are non-rotating ones and the addition of angular momentum seems to break a further half of the remaining supersymmetries.
A somewhat surprising result that deserves further study is the fact that the most general family of supersymmetric solutions of this theory (i.e. the supersymmetric Plebanski-Demianski solutions) is invariant under a continuous $`SO(2)`$ group of electric-magnetic duality transformations. Had we not included in our study NUT charge the existence of that symmetry would have passed completely unnoticed. Its meaning is, however, obscure. After all, the charges that undergo the duality rotation in its simplest, linear form, are not the physical charges. In terms of the physical charges, the duality transformations are very nonlinear.
## Acknowledgments
The work of N.A.-A., P.M. and T.O. have been supported in part by the European Union TMR program FMRX-CT96-0012 Integrability, Non-perturbative Effects, and Symmetry in Quantum Field Theory and by the Spanish grant AEN96-1655. P.M. would like to thank Iberdrola and the Universidad Autรณnoma de Madrid for their support.
## Appendix A Obtaining the Topological KN-TN-aDS Metric from PDโs
Performing in Eq. (2.11) the coordinate change (see the analogous discussion in )
$$\begin{array}{cccccc}\hfill q& =& r,\hfill & \hfill \tau & =& t+\left[a^1N^2+a\mathrm{}^2\right]\phi ,\hfill \\ & & & & & \\ \hfill p& =& N+au,\hfill & \hfill \sigma & =& a^1\varphi ,\hfill \end{array}$$
(A.1)
and the following redefinitions of the parameters $`๐ฌ=M`$, $`๐ฐ=Q`$, $`๐ฏ=P`$,
$$\begin{array}{ccc}\hfill ๐ค& =& \mathrm{}+\mathrm{}^2a^2g^2+\mathrm{\hspace{0.17em}6}g^2N^2,\hfill \\ & & \\ \hfill ๐ญ& =& N\left(\mathrm{}\mathrm{}^2a^2g^2+\mathrm{\hspace{0.17em}4}g^2N^2\right),\hfill \\ & & \\ \hfill \alpha & =& a^2\left(1+\mathrm{}\mathrm{}^2\right)N^2\left(\mathrm{}3\mathrm{}^2a^2g^2+3g^2N^2\right)+P^2,\hfill \end{array}$$
(A.2)
we go from the PD metric to the KN-TN-aDS metric as written down in Eqs. (2.7).
Note that the choice of the redefinitions is largely dictated by the factorizability of $`๐ซ`$. |
warning/0003/hep-ph0003296.html | ar5iv | text | # Fermion masses in a model for spontaneous parity breaking
## I Introduction
The origin of the left-right asymmetry in weak interactions is a longstanding problem in elementary particle physics. One possible way to understand this asymmetry is to enlarge the standard model into a left-right symmetric structure and then, by some spontaneously broken mechanism, to recover the low energy asymmetric world. Many models were developed, based on grand unified groups , superstring inspired models , a connection between parity and the strong CP problem , left-right extended standard models . All these approaches imply the existence of some new intermediate physical mass scale, well bellow the unification or the Planck mass scale. The increasing experimental evidence on neutrino oscillations and non zero masses has also motivated a renewed interest on mechanisms for parity breaking.
Left-right models starting from the gauge group $`SU(2)_LSU(2)_RU(1)_{BL}`$ were developed by many authors and are well known to be consistent with the standard $`SU(2)_LU(1)_Y`$ predictions for fundamental interactions and imply new heavy gauge bosons, not yet observed. However, for the fermion mass spectrum there is no unique choice of the Higgs sector that can reproduce the observed values for both charged and neutral fermions, neither the fundamental fermionic representation is uniquely defined.
In this paper we present a model that starts by the simple gauge structure of $`SU(2)_LSU(2)_RU(1)`$ and investigate a minimal Higgs sector that breaks the left-right symmetry. The fermion mass spectrum is studied for several choices of the Higgs sector. This paper is organized as follows: in section II we present our main assumptions for a left-right model; in section III we review the properties of new gauge interactions; in section IV we present a number of possibilities for the Higgs sector of fermion masses; in section V we show some phenomenological consequences for testing the models here proposed and in section VI we give our conclusions.
## II The model
The fundamental fermionic transformation under parity is very simple in the context of QED: it is done by the transformation between the right and left components of the same fundamental fields. Weak interactions indicates new elements in the parity transformation mechanism since the right and left components have different isospin assignments and there is no deeper understanding of this asymmetry. As the parity asymmetry is clearly displayed only in the weak isospin doublet of the fundamental fermionic representation, we take as a minimal gauge sector the left-right symmetric group $`SU(2)_LSU(2)_RU(1)_Y`$, with generators $`(T,T^{},Y)`$. The $`U(1)`$ quantum number could be gauged as B-L. This group can be considered as a sub-group of many unification groups like the superstring inspired $`E_8E_8^{}`$ or the SUSY - $`SO(10)SO(10)^{}`$. Earlier left-right models were based on the hypothesis of a symmetric world with fundamental fermions at the same mass scale of the presently known standard quarks and leptons. Some recent models propose a duplication of the full standard model gauge sector $`SU(3)SU(2)U(1)`$. More recently other models were proposed with different left and right mass scales. We consider a particular choice that avoids the introduction of a second photon, which is known to present difficulties with the positronium decay rate . We will assume this starting point and introduce new mirror fermions with the following assignment:
$$\mathrm{}_L=\left(\genfrac{}{}{0pt}{}{\nu }{e}\right)_L,\nu _R,e_R\stackrel{\mathrm{P}}{}L_R=\left(\genfrac{}{}{0pt}{}{N}{E}\right)_R,N_L,E_L$$
(II.1)
$$(\frac{1}{2},0,1);(0,0,0);(0,0,2)(0,\frac{1}{2},1);(0,0,0);(0,0,2)$$
The other lepton and quark families follow a similar pattern. This fundamental representation clearly is anomaly free. In this model the parity operation transforms the $`SU(2)_L\stackrel{\mathrm{P}}{}SU(2)_R`$ sectors, including the vector gauge bosons. This symmetry property can be used to reduce the three group constants $`g_L;g_R;g^{}`$ to only two, with $`g_L=g_R`$. For the other leptonic and quark families a similar structure is proposed. The charge generator is given by $`Q=T_3+T_3^{}+Y/2`$.
In order to break $`SU(2)_LSU(2)_RU(1)_Y`$ down to $`U(1)_{em}`$ we introduce two Higgs doublets that under parity are transformed as $`\chi _L\chi _R`$. Their quantum numbers are:
$$\chi _L\stackrel{\mathrm{P}}{}\chi _R$$
(II.2)
$$(\frac{1}{2},0,1)(0,\frac{1}{2},1)$$
and the corresponding symmetry breaking is realized through the vacuum parameters $`v_L`$ and $`v_R`$. The model includes a Higgs field $`\mathrm{\Phi }`$ in the mixed representation (1/2,1/2,0). The symmetry breaking is done by the neutral components, with parameters $`k`$ and $`k^{}`$. Finally we consider scalar fields in the representation (0,0,0), broken at $`s_{GUT}`$ or at lower scales. In the standard model, and also in some of its extentions, Higgs doublets are responsible for the gauge boson masses as well as for the fermion masses. For the gauge bosons this hypothesis is strongly confirmed by the experimental result $`\rho =1`$. For the fermion masses the standard model requires adjusting by hand the Yukawa couplings in order to reproduce the observed mass spectrum. Although this procedure is consistent in the sense that all couplings satisfy $`g_i<1`$, there is no direct experimental confirmation for this hypothesis. One of the main points of our work is to show that Higgs singlets in left-right models can give a consistent charged and neutral fermion mass spectrum. We are also supposing the symmetry breaking hierarchy $`s_{GUT}>>v_R>>v_L,k,k^{}`$.
## III The gauge bosons
In order to make our presentation more complete we briefly review in this section the main properties of the gauge sector in the $`SU(2)_LSU(2)_RU(1)_Y`$ model.
The symmetry breaking mechanism follows the standard model procedure and, with the above hierarchy, we readily find for the charged vector bosons the masses (for $`v_L>>k,k^{}`$):
$`M_{W_L}^2`$ $``$ $`{\displaystyle \frac{1}{4}}g_L^2v_L^2`$ (III.1)
$`M_{W_R}^2`$ $``$ $`{\displaystyle \frac{1}{4}}g_R^2v_R^2`$ (III.2)
We call attention to the fact that the above Higgs field in the mixed representation (1/2,1/2,0) implies a mixing between $`W_L`$ and $`W_R`$. This mixing is given by $`\mathrm{sin}\theta _M=kk^{}/v_R^2`$. It is well known that this mixing is strongly suppressed, as required by the absence of right-handed currents . From the first equation III.1, we have the identification $`v_L=v_{Fermi}`$ and the charged current interactions of standard fermions must be according to $`(1/2;0;Y)W_L`$ and for the new mirror fermions $`(0;1/2;Y)W_R`$. There are alternatives to this scheme. We can also have $`v_L=0`$ and the Fermi scale is given by $`v_{Fermi}^2k^2+k^2`$ with a small mixing between $`W_L`$ and $`W_R`$.
Before proceeding to the neutral sector it is convenient to introduce the notation,
$`sin^2\theta _W`$ $``$ $`{\displaystyle \frac{g_R^2g^2}{g_R^2g_L^2+g_R^2g^2+g_L^2g^2}}`$ (III.3)
$`sin^2\beta `$ $``$ $`{\displaystyle \frac{g^2}{g_R^2+g^2}}`$ (III.4)
This is a change from ($`g_L`$; $`g_R`$; $`g^{}`$) to the basis $`(g_L;\mathrm{sin}\theta _W;\mathrm{sin}\beta )`$. The condition $`g_L=g_R`$ implies $`\mathrm{sin}\beta =\mathrm{tan}\theta _W`$. We also introduce the ratio $`\omega =v_L/v_R`$, which is supposed to satisfy $`\omega ^21`$ due to the non observation of any new physical scale with energies up to 1 TeV.
For the neutral vector boson sector we find, after diagonalization and expanding in powers of $`\omega `$,
$`M_\gamma `$ $``$ $`0`$ (III.5)
$`M_Z^2`$ $``$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{v_L^2g_L^2}{\mathrm{cos}^2\theta _W}}\{1\omega ^2\mathrm{sin}^4\beta \}`$ (III.6)
$`M_Z^{}^2`$ $``$ $`{\displaystyle \frac{1}{4}}v_R^2g_L^2\mathrm{tan}^2\theta _W\mathrm{tan}^2\beta \{1+{\displaystyle \frac{\omega ^2sin^22\beta }{4\mathrm{sin}^2\theta _W}}\}`$ (III.7)
The unification condition for the electromagnetic interaction is the same as in the standard model,
$$e=g_L\mathrm{sin}\theta _W,$$
(III.8)
and the $`\rho `$ parameter is changed to
$$\rho =\frac{M_{WL}^2}{M_Z^2\mathrm{cos}^2\theta _W}\{1\omega ^2\mathrm{sin}^4\beta \}$$
(III.9)
In order to compare the model with the experimental data, we must develop the neutral current interactions. First we recall that the gauge sector of the model has five input parameters $`g_L`$; $`g_R`$; $`g^{}`$; $`v_L`$; $`v_R`$. The condition $`g_L=g_R`$ reduces them to four. We take as experimental inputs $`M_W=80.4`$ GeV; $`\mathrm{sin}^2\theta _W=0.2230`$ and $`\alpha (M_Z)=1/128`$. The W mass is much less precise than the Z mass, but we have taken this procedure since from equation III.1 the determination of the fundamental parameters is much more transparent. If we start from the Z mass there is no significative changes in our conclusions. With the values above we readily obtain $`v_L=248`$ GeV and $`g_L=0.6635`$ as in the standard model and
$$\mathrm{sin}^2\beta =0.286$$
(III.10)
Then the only unknown in the gauge sector of the model is $`v_R`$ which is supposed to be higher than $`v_L`$.
The neutral currents coupled to the massive vector bosons Z and Zโ are given by:
$`J_\mu `$ $`=`$ $`{\displaystyle \frac{g_L}{\mathrm{cos}\theta _W}}\gamma _\mu \{(1\omega ^2\mathrm{sin}^4\beta )T_3\omega ^2\mathrm{sin}^2\beta \mathrm{cos}^2\beta T_3^{}`$ (III.11)
$``$ $`Q\mathrm{sin}^2\theta _W(1{\displaystyle \frac{\omega ^2\mathrm{sin}^4\beta }{\mathrm{sin}^2\theta _W}})\}`$ (III.12)
$`J_\mu ^{}`$ $`=`$ $`g_L\mathrm{tan}\theta _W\mathrm{tan}\beta \gamma _\mu \{(1+{\displaystyle \frac{\omega ^2\mathrm{sin}^2\beta \mathrm{cos}^2\beta }{\mathrm{sin}^2\theta _W}})T_3`$ (III.13)
$`+`$ $`{\displaystyle \frac{1}{\mathrm{sin}^2\beta }}T_3^{}Q(1+\omega ^2\mathrm{cos}^2\beta \mathrm{sin}^2\beta )\}`$ (III.14)
From the first of the preceding equations we clearly recover the standard model result for $`J_\mu (Z)`$ in the limit $`\omega ^20`$. The new neutral fermionโs coupling with the standard Z is suppressed by a factor $`\omega ^2`$. However, the standard fermionโs couplings with the new Zโ are not suppressed, according to the second of equations III.7. The main question at this point is how small $`\omega `$ can be in order that the model agrees with present data. It is well known that the presently known high precision data requires the inclusion of the standard model quantum corrections, in order to make a consistent comparison between theory and experiment. To compare the model with data we have considered that the new terms in equations III.7 are small corrections to the standard model predictions and at least of the same order of the quantum corrections. We have computed the corrections to the Z couplings to standard fermions and found
$`g_V^{L,R}`$ $`=`$ $`g_V^{SM}(\omega ^2\mathrm{sin}^4\beta )\{T_3^L2Q\}`$ (III.15)
$`g_A^{L,R}`$ $`=`$ $`g_A^{SM}(\omega ^2\mathrm{sin}^4\beta )\{T_3^L\}`$ (III.16)
The Particle Data Group, in their 2000 edition , summarizes the present data from low energy lepton interaction, lepton-hadron collisions and the high precision data from LEP and SLAC. They also present the experimental averages for the $`g_V`$ and $`g_A`$ couplings for charged and neutral leptons. The most stringent bounds comes from the effective coupling of the Z to the electron neutrino $`g_{exp}^{\nu e}=0.528\pm 0.085`$ and $`\mathrm{\Gamma }_{exp}^{inv}(Z)=498.8\pm 1.5`$ MeV, to be compared with the standard model predictions $`g_{SM}=0.5042`$ and $`\mathrm{\Gamma }_{SM}^{inv}(Z)=501.65\pm 0.15`$ MeV. For the muon neutrino coupling with the Z boson, the Particle Data Group quotes $`g_{exp}^{\nu \mu }=0.502\pm 0.017`$. We have performed a fit to this data, using the standard model predictions, and we find that deviations from the standard model must be bounded, at $`95\%`$ confidence level by:
$$(\omega ^2\mathrm{sin}^4\beta )<10^4$$
(III.18)
This bound is consistent with the present experimental constraint on the $`\rho `$ parameter. With the value for $`\mathrm{sin}\beta `$ given in equation III.6, we have the bound
$$v_R>30v_L$$
(III.19)
For the new Zโ mass we have
$$M_Z^{}>800GeV$$
(III.20)
This value is a little above the present experimental bounds on new gauge bosons searches done by the CDF and D0 collaborations at Fermilab.
## IV The fermion masses
The fermion mass spectrum depends both on the Higgs choice of the model and on the fundamental fermionic representation. A particular property of the model is the presence of left and right handed singlets in the fundamental representation. This means that we can add to the mass lagrangian new bare terms or new Higgs scalars which have no consequences on the vector gauge boson masses. We consider two new Higgs singlets, one that is coupled to Dirac terms in the mass lagrangian - $`S_D`$ \- and the other that couples to Majorana terms - $`S_M`$. After spontaneous symmetry breaking they develop vacuum parameters $`s_D`$ and $`s_M`$, respectively.
The most general Yukawa lagrangian is given by,
$``$ $`=`$ $`f\{\overline{\mathrm{}_L}\stackrel{~}{\chi }_L\nu _R+\overline{L_R}\stackrel{~}{\chi }_RN_L\}+f^{}\{\overline{\mathrm{}_L}\stackrel{~}{\chi }_LN_L^C+\overline{L_R}\stackrel{~}{\chi }_R\nu _R^C\}`$ (IV.1)
$`+`$ $`gS_M\{\overline{N_L^C}N_L+\overline{\nu _R^C}\nu _R\}+g^{}S_D\{\overline{\nu _R}N_L\}`$ (IV.2)
$`+`$ $`f\mathrm{"}\{\overline{\mathrm{}_L}\mathrm{\Phi }L_R\}+g^{}S_D\{\overline{e_R}E_L\}+h.c.`$ (IV.3)
For the neutral fermions, the mass lagrangian after symmetry breaking is given by
$``$ $`=`$ $`fv_L\overline{\nu _L}\nu _R+f^{}v_L\overline{\nu _L^c}N_L+fv_R\overline{N_R}N_L+f^{}v_R\overline{N_R^C}\nu _R+`$ (IV.4)
$`+`$ $`gs_M\overline{N_L^C}N_L+gs_M\overline{\nu _R^C}\nu _R+`$ (IV.5)
$`+`$ $`g^{}s_D\overline{\nu _R}N_L+g\mathrm{"}k\overline{\nu _L}N_R,`$ (IV.6)
For the charged fermions we have a similar lagrangian, except for the conjugated terms. The generalization for the other families is straightforward. The diagonalization is most easily done by introducing the self conjugated fields
$`\chi _i`$ $`=`$ $`\psi _{iL}+\psi _{iL}^C`$ (IV.7)
$`\omega _j`$ $`=`$ $`\psi _{jR}+\psi _{jR}^C`$ (IV.8)
with $`i,j=\nu ,N`$.
We will consider in this paper that the order of magnitude of the fermionic mass spectrum is given by the symmetry breaking scales and their combinations. This is a departure from the standard model procedure of adjusting the coupling constants in the Yukawa interactions. So we are supposing that all coupling constants are of order one.
In the basis $`(\chi _\nu ;\omega _N;\chi _N;\omega _\nu )`$ the general neutrino mass matrix is:
$$M_{\nu ,N}=\left(\begin{array}{cccc}0& \frac{k}{2}& v_L& \frac{v_L}{2}\\ \frac{k}{2}& 0& \frac{v_R}{2}& v_R\\ v_L& \frac{v_R}{2}& s_M& \frac{s_D}{2}\\ \frac{v_L}{2}& v_R& \frac{s_D}{2}& s_M\end{array}\right)$$
and the charged fermion mass matrix is:
$$M_{e,E}=\left(\begin{array}{cccc}0& k^{}& 0& v_L\\ k^{}& 0& v_R& 0\\ 0& v_R& 0& s_D\\ v_L& 0& s_D& 0\end{array}\right)$$
For this last case, we recover the Dirac formalism by the standard $`\pi /4`$ rotations over the Majorana fields.
After the results of the Super Kamiokande collaboration there are strong experimental evidence that neutrinos oscillate and have non-zero mass . Atmospheric neutrino experiments are consistent with $`\nu _\mu \nu _\tau `$ oscillations with a large mixing and $`\mathrm{\Delta }m^2=(1.56)10^3`$ $`eV^2`$. Solar neutrino evidence still allow some alternative solutions like vacuum oscillations and large and small angle MSW mechanism. If one considers the LSND results then the solution for the neutrino parameters seems to indicate the need of a forth sterile neutrino, but so far there is no confirmation of this result. In phenomenological models for quark and lepton masses we have a number of possibilities for fermion mass textures in models such as $`SO(10)`$ . In view of the present experimental situation we will not proceed to fit all the neutrino masses and mixings but we will look for solutions which could accommodate neutrinos with masses in the $`10^210^3`$ eV range. We present three possible solutions, which differ in the choice of the symmetry breaking parameters.
$``$ Model I
In this model we consider that the mixed Higgs field do not break the vacuum symmetry ($`k=k^{}=0`$) and that the Higgs singlets are both broken at $`s_{GUT}`$. For neutrinos we have the following masses
$`m_{\nu 1}`$ $`=`$ $`v_L^2/s_{GUT}`$ (IV.9)
$`m_{\nu 2}`$ $`=`$ $`v_R^2/s_{GUT}`$ (IV.10)
$`m_{N1}`$ $``$ $`m_{N2}s_{GUT}`$ (IV.11)
and for the charged fermions
$`m_e`$ $`=`$ $`v_Lv_R/s_{GUT}`$ (IV.13)
$`m_E`$ $`=`$ $`s_{GUT}`$ (IV.14)
In this model we have the โuniversal see-sawโ mass relation $`m_{\nu 1}m_{\nu 2}=m_e^2`$. Using $`v_L=v_{Fermi}`$ and $`s_{GUT}=10^{16}`$ GeV, we must have $`v_R=10^{10}`$ in order to obtain the correct value for the electron mass. The first generation mass spectrum is then given by $`m_{\nu 1}10^2`$ eV ; $`m_{\nu 2}10`$ TeV ; $`m_{N1}m_{N2}10^{16}`$ GeV; $`m_e1`$ MeV; $`M_E10^{16}`$ GeV. The smallness of the electron mass is a consequence of a โsee-sawโ mass relation given by equation IV.5. This is a departure for the standard model mechanism for fermion masses. There is no mixing between $`\nu _1`$ and $`\nu _2`$ and we have an example of a sterile neutrino coming from a new exotic doublet. The presently observed neutrino mixing must come from a possible generation mixing in the general neutrino mass lagrangian. Standard charged quarks are also found to be in the MeV mass range, with $`m_q=v_Lv_R/s_{GUT}`$. With this high value for $`v_R`$ we have no experimental accessible accelerator possibilities for new gauge vector bosons. For the Large Hadron Collider (LHC) it has recently been shown that heavy neutrino production is limited to masses of a few hundreds of GeV. Heavy neutrinos with masses in the TeV region can be produced in the next generation of $`e^+e^{}`$ or $`e^{}\mu ^+`$ colliders.
The $`v_R`$ value consistent with the electron mass is of the order of the Peccei-Quinn symmetry breaking scale and we can have a possible explanation of the small value for the $`\theta `$-angle of the strong CP problem, as shown in references . In order to generate the other families mass spectrum we have two possibilities. The first one, following the standard model procedure, is simply to adjust the coupling constants in the general mass lagrangians. The other possibility is to enlarge the Higgs singlet sector, postulating one new field for each family. In this last case we have an hierarchy for neutrino masses.
$``$ Model II
In this model we impose explicit lepton number conservation and $`k=k^{}=0`$.The Majorana mass terms must be zero, except for the Higgs singlet coupled to Majorana mass terms in the neutrino sector that can have two units of leptonic number. Lepton number is then spontaneously broken by this term at a scale $`s_Ms_{GUT}`$ and the Higgs singlet coupled to Dirac mass terms is allowed to be broken at a lower scale. The neutrino mass spectrum is given by
$`m_{\nu 1}`$ $`=`$ $`v_L^2/s_M`$ (IV.15)
$`m_{\nu 2}`$ $`=`$ $`v_R^2/s_M`$ (IV.16)
$`m_{N_1,N_2}`$ $`=`$ $`s_M\pm s_D/2`$ (IV.17)
and for the charged fermions
$`m_e`$ $`=`$ $`v_Lv_R/s_D`$ (IV.19)
$`m_E`$ $`=`$ $`s_D`$ (IV.20)
If we take $`v_L=v_{Fermi}`$ and $`s_M=s_{GUT}`$, then the electron and quark masses allow a solution given by $`v_R10^4`$ GeV and $`s_D10^{10}`$ GeV. Here again we have a โsee-sawโ mechanism for the electron mass. In this model we could have an experimentally accessible new neutral current, as shown in Fig. 1. One light neutrino has a mass in the $`10^210^3`$ eV range and the other is in the $`10.1`$ eV region. These neutrinos are orthogonal and again we have a new sterile neutrino. The Peccei-Quinn symmetry breaking scale reappears in the Higgs singlet sector. Family replication can be recovered with different Dirac singlets and for only one Majorana mass scale we can have a degenerate neutrino mass spectrum.
$``$ Model III
We consider the same Higgs sector as in Model I, except for the mixed representation, with a small value for $`k^{}`$. The electron mass is changed to $`m_e=k^{}+v_Lv_R/s_{GUT}`$. If we adjust $`k^{}=m_e`$ then there is no need to a large $`v_R`$ as in the first model. The down-quark mass is given by the same value of $`k^{}`$. However, for the up-quark mass we need a new $`kk^{}`$ in the mixed representation and this will not give a correct neutrino mass. An alternative to this difficulty is to consider that neutrinos are massless at tree level and acquire masses through quantum corrections . As this mechanism involves new additional hypothesis we will not develop the full solution here. We can recover the bound $`v_R=10^310^4`$ GeV and we have the possibility to test this model through the new $`Z^{}`$ interactions.
The generalization for the other families can be easily extended from the above arguments. However, the mixing angle pattern is not so simple. A first approach is to generalize the see-saw mechanism, with mixing angles given by the mass ratios $`\theta _{mix}m_\nu /m_N`$, which are very small numbers. There are many models that avoid such a restriction: the introduction of an arbitrary number of right-handed neutrinos ; some fine-tuning in the neutrino mass matrix and any general singular neutrino mass matrix can disconnect mixing parameters from mass ratios. So we will take mixing angles and neutrino masses as independent parameters. The high precision experimental data from LEP and SLC gives the bound $`\mathrm{sin}\theta _{mix}^2<10^210^3`$.
## V Phenomenological consequences
Let us now turn our attention to some of the phenomenological consequences of the models developed in the preceding section. For Model I the high value of $`v_R`$ makes new gauge bosons out of reach even for the next generation of accelerators. The model can be tested via the new mirror fermions, coupled to the standard model neutral current according to equation III.7. New charged leptons are coupled to the Z proportional to $`\mathrm{sin}^2\theta _W`$ and their phenomenology was extensively studied by many authors . The new neutral heavy leptonโs coupling to Z is suppressed by a factor $`\omega ^2\mathrm{sin}^4\beta <10^4`$ and can be detected only if heavy neutrinos are mixed with light neutrinos. This introduces a new mixing parameter that it is also known to be small , of the order of $`\alpha _{mix}^210^210^3`$. The $`N\overline{N}`$ coupling to Z is suppressed by $`\mathrm{sin}^2\alpha _{mix}`$ and the light-to-heavy neutrino coupling to Z is given by a single power of $`\mathrm{sin}\alpha _{mix}`$. This means that single heavy neutrino production is favored relative to pair production .
In Fig. 1 we show the total cross section for pair and single Dirac heavy neutrino production in $`e^+e^{}`$ collisions at a new Next Linear Collider (NLC) at $`\sqrt{s}=2`$ TeV. For a single heavy neutrino production we can identify the signal by itโs decay in $`e^\pm W^{}`$. These processes can be readily calculated by using high energy algebraic programs such as CompHep .
In Fig. 2 we show the $`eW`$ invariant mass distribution for both the signal and standard model background for $`e^+e^{}e^\pm \nu _eW^{}`$ and $`M_{Ne}=`$ 400 GeV. In this process we have 6 diagrams from the new heavy lepton and 12 from the standard model background . Here again we can separate the signal from the standard model background, if we take the upper bound $`\mathrm{sin}^2\alpha _{mix}<0.0052`$ .
In Model II we can have the production of new heavy neutral gauge bosons with mass scales accessible at hadron colliders. Dilepton production in hadron-hadron collisions give a very clear signal for a new Zโ. The new facilities at the Tevatron proton-antiproton collider, with $`\sqrt{s}=2`$ TeV and a luminosity of 1000 pb<sup>-1</sup>, will allow the search of the new Zโ with masses in the 800-1000 GeV region. For $`p\overline{p}Z^{}\mathrm{}^\pm \mathrm{}^{}`$, with $`\mathrm{}=e,\mu `$, we expect 20 events for a Zโ mass of 800 GeV. We have employed the parton distribution functions CTEQ4m by .
The Large Hadron Collider (LHC) facilities using proton-proton collisions at $`\sqrt{s}=14`$ TeV, will attain higher Zโ masses in the 1-4 TeV region. We show in Fig. 3 the invariant mass distribution for the $`Z^{}`$ decay at the LHC, with $`M_Z^{}=2`$ TeV in the leptonic channel $`Z^{}\mathrm{}^\pm \mathrm{}^{}`$ with $`\mathrm{}=e,\mu `$. For an integrated luminosity at LHC of 100 fb<sup>-1</sup> and cuts of $`E_{\mathrm{}}>20`$ GeV and $`|\eta |<2,5`$ we expect 1000 events for $`M_Z^{}=1`$ TeV and only one event for $`M_Z^{}=4`$ TeV. New heavy neutrinos can also be coupled to new heavy neutral gauge bosons and can be produced at the new large hadron collider. A detailed study of these possibilities was done in Ref. .
## VI Conclusions
In conclusion we have shown that spontaneously broken parity models with a consistent fermion mass spectrum can be built at intermediate scales ranging from $`10^3`$ to $`10^{10}`$ GeV. New mirror fermions are present and can be connected to ordinary fermions through neutral currents. We propose a simple Higgs sector for the model that allows several physically interesting solutions. The charged fermion masses can be generated by a โsee-sawโ mass relation, analogous to the neutrino sector. Various scenarios for neutrino masses are possible and a more clear experimental definition on neutrino masses will allow to test their reality. Neutral heavy gauge bosons Z and Zโ, coupled to ordinary and new mirror fermions, are expected to play a fundamental role in the understanding of the left-right symmetry. This is a departure from other models that propose gravitation as the connection between the left and right sectors. Experimental consequences of the model could be found at the next generation of colliders.
Acknowledgments: This work was partially supported by the following Brazilian agencies: CNPq, FUJB, FAPERJ and FINEP.
Figure Captions
1. Total cross section for pair and single heavy neutrino production for $`e^+e^{}`$ collisions at NLC collider with $`\sqrt{s}=2`$ TeV. $`N_e`$ corresponds to a heavy neutrino of the electron family.
2. Invariant $`eW`$ mass distribution in the process $`e^+e^{}e^+\nu _eW^{}`$ at NLC with $`\sqrt{s}=2`$ TeV and $`M_{Ne}=400`$ GeV. The flat part of the curve is the standard model background.
3. The invariant mass distribution for $`ppZ^{}\mathrm{}^\pm \mathrm{}^{}`$ (where $`\mathrm{}=`$e or $`\mu `$) for pp collisions at LHC with $`\sqrt{s}=14`$ TeV, $`E>20`$ GeV and $`|\eta |<2,5`$ and $`M_Z^{}=2`$ TeV. |
warning/0003/gr-qc0003070.html | ar5iv | text | # Vector field mediated models of dynamical light velocity
## 1. Introduction
In previous work , we have introduced a new kind of vector-tensor and scalar-tensor theory of gravity, which exhibits a bimetric structure and contains two or more light cones. This type of model has attracted some interest recently , and similar effects have been noted elsewhere . The motivation for considering these models is derived form earlier work of one of the present authors , which provided a scenario in which some of the outstanding issues in cosmology can be resolved. The present line of work provides a specific class of models that realize these ideas, for it provides a fundamental dynamical mechanism for varying speed of light theories and generates a new mechanism for an inflationary epoch that could solve the initial value problems of early universe cosmology.
In this article, we focus on clarifying the role that these models can play in the early universe, demonstrating how matter that satisfies the strong energy condition can nevertheless contribute to the cosmic acceleration. Recently, an analysis of a similar class of theories has appeared which, while introducing some interesting ideas, unfortunately claimed that we had made an algebraic error in our previous work . This can be attributed to a misunderstanding of part of the construction that was perhaps not dealt with in sufficient detail in our initial publication. We will rectify this situation here, developing the model in additional detail and showing explicitly that the error attributed to us in is in fact an error on their part. We will also show that our cosmological model can be mapped to a model with varying fundamental constants , albeit not uniquely and requiring some care in the interpretation of the varying constants that appear.
It is hoped that the models can shed some light on the new observational data that suggests the expansion of the universe at present is undergoing an acceleration . Although there has been some success in understanding the latter problem by the inclusion of a class of very particular scalar field potentials , it is fair to say that not all issues have been resolved. In this article, we will not have much to say about this issue since, as we shall see, the vector field that produces a superluminary expansion in the early universe must vanish at some scale, and standard cosmology results afterwards. Using the scalar field version of the model, we expect that not only will we be able to generate sufficient inflation, but that a quintessence-like solution should be achieveable. We shall concentrate our efforts on the vector-tensor model by providing a more detailed analysis of its consequences and postpone a fuller analysis of the scalar-tensor model to a future article.
## 2. The vector mediated model
We shall be considering models wth an action of the form
$$S=\overline{S}_{\mathrm{gr}}[\overline{\mathrm{g}}]+S[\mathrm{g},\psi ]+\widehat{S}[\widehat{\mathrm{g}},\widehat{\varphi }^I].$$
(1)
The first term is the usual Einstein-Hilbert action for general relativity constructed from a metric $`\overline{\mathrm{g}}_{\mu \nu }`$, and the final term is the contribution from the non-gravitational (matter) fields in spacetime $`\widehat{\varphi }^I`$, and is built from a different but related metric $`\widehat{\mathrm{g}}_{\mu \nu }`$. One motivation for constructing the action (1) is simplicity: it allows us to build models in a modular way with additional matter fields introduced as necessary, and it requires only fairly well-known variational results. The other benefit is that it makes fairly clear what metrics will be of physical relevance.
The contribution $`S[\mathrm{g},\psi ]`$ is constructed from a metric $`\mathrm{g}_{\mu \nu }`$ and includes kinetic terms for a field or fields (unspecified as yet) $`\psi `$ that may be considered to be part of the gravitational sector, modifying the reaction of spacetime to the presence of the matter fields in $`\widehat{S}[\widehat{\mathrm{g}},\widehat{\varphi }^I]`$. The manner in which that $`\psi `$ accomplishes this is by modifying the metric that appears in each of the actions. For example, in $`\psi `$ was a vector field, $`\overline{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }`$ and $`\widehat{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }+b\psi _\mu \psi _\nu `$, whereas in $`\psi `$ was a scalar field, $`\overline{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }`$ and $`\widehat{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }+b_\mu \psi _\nu \psi `$. These relations imply that matter and gravitational fields propagate at different velocities if $`\psi `$ is nonvanishing. In this work we will explore the first of these possibilities in more detail than was possible in our previous publication , clarifying some misinterpretations that have appeared in recent work .
Since the matter action $`\widehat{S}`$ is built using only $`\widehat{\mathrm{g}}_{\mu \nu }`$, it is the null surfaces of $`\widehat{\mathrm{g}}_{\mu \nu }`$ along which matter fields propagate. If we assume that other than the presence of a โcompositeโ metric the matter action is otherwise a conventional form (perfect fluid, scalar field, Maxwell, etc.), then variation of the matter action
$$\delta \widehat{S}[\widehat{g},\widehat{\varphi }^I]=๐\widehat{\mu }\left(\widehat{F}_I\delta \widehat{\varphi }^I\frac{1}{2}\widehat{T}^{\mu \nu }\delta \widehat{\mathrm{g}}_{\mu \nu }\right),$$
(2)
provides the matter energy-momentum tensor $`\widehat{T}^{\mu \nu }`$, which will be conserved
$$\widehat{}_\nu \widehat{T}^{\mu \nu }=0,$$
(3)
by virtue of the matter field equations $`\widehat{F}_I=0`$. Throughout we will write, for example, $`\widehat{}_\nu `$ for the covariant derivative constructed from the Levi-Civita connection of $`\widehat{\mathrm{g}}_{\mu \nu }`$. Since we also assume that the matter fields satisfy the dominant energy condition, we therefore know (assuming appropriate smoothness of $`\widehat{\mathrm{g}}_{\mu \nu }`$) that matter fields cannot travel faster than the speed of light as determined by $`\widehat{\mathrm{g}}_{\mu \nu }`$ .
The gravitational action is written
$$\overline{S}_{\mathrm{gr}}[\overline{\mathrm{g}}]=\frac{1}{\kappa }๐\overline{\mu }\overline{R},$$
(4)
where we use a metric with ($`+`$$``$$``$$``$) signature and have defined $`\kappa =16\pi G/c^4`$. We will denote the metric densities by, e.g., $`\overline{\mu }=\sqrt{det(\overline{\mathrm{g}}_{\mu \nu })}`$ and in addition write $`d\overline{\mu }=\overline{\mu }dtd^3x`$. The variation of (4) is
$$\delta \overline{S}_{\mathrm{gr}}[\overline{\mathrm{g}}]=\frac{1}{\kappa }๐\overline{\mu }\overline{G}^{\mu \nu }\delta \overline{\mathrm{g}}_{\mu \nu }.$$
(5)
We will not consider a cosmological constantโit is a trivial matter to include it later. Provided the resulting field equation for $`\overline{G}^{\mu \nu }`$ has nothing in it that disturbs the principal part of the field equations and the constraints remain constraints, then we can identify the metric $`\overline{\mathrm{g}}_{\mu \nu }`$ as providing the light cone for the gravitational system.
It remains to connect these two pieces with specific models for $`S`$, $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$. We will generalize and provide more details on the vector field model that we presented in . The choice of this model over a scalar field mediated model is due to the fact that the vector field models do not have any remaining degrees of freedom in an FRW universe and are therefore simpler to analyze.
Keeping in mind the dangers involved in coupling vector fields to gravity , we begin with a Proca model with arbitrary potential
$$S[\mathrm{g},\psi ]=\frac{1}{\kappa }๐\mu \left(\frac{1}{4}B^2V(X)\right),$$
(6)
where we will use the definition
$$X=\frac{1}{2}\psi ^2,$$
(7)
and $`V^{}(X)=V(X)/X`$. We will also use $`B_{\mu \nu }=_\mu \psi _\nu _\nu \psi _\mu `$, $`\psi ^2=\mathrm{g}^{\mu \nu }\psi _\mu \psi _\nu `$ and $`B^2=\mathrm{g}^{\alpha \mu }\mathrm{g}^{\beta \nu }B_{\alpha \beta }B_{\mu \nu }`$. We will assume that as $`\psi _\mu 0`$ we have $`V(X)m^2X`$ and therefore the linearized (in $`\psi _\mu `$) limit of our model is identical to Einstein-Proca field equations coupled to matter.
Using the standard energy-momentum tensor for the vector field
$$T^{\mu \nu }=B^{\mu \alpha }B_{}^{\nu }{}_{\alpha }{}^{}+\frac{1}{4}\mathrm{g}^{\mu \nu }B^2+V^{}\psi ^\mu \psi ^\nu V\mathrm{g}^{\mu \nu },$$
(8)
the variation of (6) results in
$$\delta S[\mathrm{g},\psi ]=\frac{1}{\kappa }๐\mu \left(_\mu B^{\mu \alpha }+V^{}\psi ^\alpha \right)\delta \psi _\alpha \frac{1}{2\kappa }๐\mu T^{\mu \nu }\delta \mathrm{g}_{\mu \nu }.$$
(9)
Note that writing $`\psi ^\alpha `$ is potentially ambiguous, since there is more than one way in which one can โraiseโ an index. We will use $`\psi ^\alpha =\mathrm{g}^{\alpha \beta }\psi _\beta `$, $`\widehat{\psi }^\alpha =\widehat{\mathrm{g}}^{\alpha \beta }\psi _\beta `$ and $`\overline{\psi }^\alpha =\overline{\mathrm{g}}^{\alpha \beta }\psi _\beta `$ as necessary, and consider $`\psi _\mu `$ as the independent components of the (co-)vector field. The same will hold for any other tensor, and we will be explicit about which contraction we are using where necessary.
In choosing the form of $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$, we could be fairly general and define, for example
$$\widehat{\mathrm{g}}_{\mu \nu }=a(B^2,X)\mathrm{g}_{\mu \nu }+b(B^2,X)\psi _\mu \psi _\nu +\text{other terms like }B_{}^{\alpha }{}_{\mu }{}^{}B_{\alpha \nu }\text{etc.}.$$
(10)
The presence of $`B_{\mu \nu }`$ in this relation would lead to second derivatives of $`\psi _\mu `$ appearing in the relationship between $`\mathrm{\Gamma }_{\mu \nu }^\alpha `$ and $`\widehat{\mathrm{\Gamma }}_{\mu \nu }^\alpha `$, and therefore re-writing the field equations in terms of $`\widehat{\mathrm{g}}_{\mu \nu }`$ will have a nontrivial effect on the principal parts of the differential equations. What one needs is for $`_\mu `$ and $`\widehat{}_\mu `$ to differ only by first derivatives of $`\psi _\mu `$, which is accomplished by considering the simpler form $`\widehat{\mathrm{g}}_{\mu \nu }=a(X)(\mathrm{g}_{\mu \nu }+b(X)\psi _\mu \psi _\nu )`$. Although this more general class of models is worth considering, here we will limit ourselves to the choice
$$\widehat{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }+b\psi _\mu \psi _\nu ,\overline{\mathrm{g}}_{\mu \nu }=\mathrm{g}_{\mu \nu }+g\psi _\mu \psi _\nu ,$$
(11)
so that the variations of $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$ are related to those of $`\mathrm{g}_{\mu \nu }`$ and $`\psi _\mu `$ by, for example
$$\delta \widehat{\mathrm{g}}_{\mu \nu }=\delta \mathrm{g}_{\mu \nu }+b(\psi _\mu \delta \psi _\nu +\psi _\nu \delta \psi _\mu ).$$
(12)
This type of model is therefore a โpurely dynamical metric theoryโ, for all fields are determined by the coupled field equations, but the relationship (11) must be considered as โprior geometryโ . The field equations
$$_\mu B^{\mu \nu }+V^{}\psi ^\nu +gT^{\mu \nu }\psi _\mu +\kappa \frac{\widehat{\mu }}{\mu }(gb)\widehat{T}^{\nu \mu }\psi _\mu =0,$$
(13a)
$$\overline{\mu }\overline{G}^{\mu \nu }=\frac{1}{2}\mu T^{\mu \nu }+\frac{1}{2}\kappa \widehat{\mu }\widehat{T}^{\mu \nu },$$
(13b)
result from assembling the contributions from (2), (5) and (9) proportional to $`\delta \psi _\nu `$ and $`\delta \mathrm{g}_{\mu \nu }`$, respectively.
As we shall see from (15), making the local redefinition of metric fields (11) will not alter the principal part of any second-order field equation. We are therefore free to write the entire system (13) as a partial differential equation in terms of the fields $`\overline{\mathrm{g}}_{\mu \nu }`$, $`\psi _\mu `$ and $`\widehat{\varphi }^I`$. While it is possible that constraints in the matter system $`\widehat{F}_I=0`$ will pick up terms containing first derivatives of $`\psi _\mu `$, they will nonetheless remain constraints. It is therefore clear that $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$ provide the characteristic surfaces for matter and gravitational fields, respectively. Unfortunately, although most of the construction of our model relies on the metric $`\mathrm{g}_{\mu \nu }`$ and its variation in the action, it is not actually of any obvious physical relevance. This can be seen by considering the divergence of (13a) given in (18). When this is used to write the principal part of (13a) as a simple wave operator, we find that the derivatives of $`\psi _\mu `$ appearing in (18) make this impossible. The analysis of the characteristic speeds of such a wave operator are far from trivial (see, however, for the speeds of cosmological perturbations).
## 3. The Bianchi identities
A potential concern when considering the field equations (13b) is how the Bianchi identities $`\overline{}_\nu \overline{G}^{\mu \nu }=0`$ and matter conservation laws (3) combine to ensure that the divergence of (13b) does not lead to any additional constraints on the system. Although it could be said that covariance guarantees this consistency, because we are not considering an explicit matter model, it is important to show the relationship between any matter model satisfying (3) and the Bianchi identities.
From the definitions (11) we find, for example
$$\widehat{\mathrm{g}}_{\mu \nu }:=\mathrm{g}_{\mu \nu }+b\psi _\mu \psi _\nu ,\widehat{\mathrm{g}}^{\mu \nu }=\mathrm{g}^{\mu \nu }\frac{b}{1+2bX}\psi ^\mu \psi ^\nu ,$$
(14)
with similar relationships between the metrics $`\mathrm{g}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$ and the metrics $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$, with suitable replacement of the constant $`b`$ and re-definition of $`X`$. It is then a straightforward exercise to find
$`\widehat{\mathrm{\Gamma }}_{\mu \nu }^\alpha \mathrm{\Gamma }_{\mu \nu }^\alpha `$ $`={\displaystyle \frac{b}{1+2bX}}\psi ^\alpha _{(\mu }\psi _{\nu )}b\widehat{\mathrm{g}}^{\alpha \beta }B_{\beta (\mu }\psi _{\nu )},`$ (15a)
$`\overline{\mathrm{\Gamma }}_{\mu \nu }^\alpha \mathrm{\Gamma }_{\mu \nu }^\alpha `$ $`={\displaystyle \frac{g}{1+2gX}}\psi ^\alpha _{(\mu }\psi _{\nu )}g\overline{\mathrm{g}}^{\alpha \beta }B_{\beta (\mu }\psi _{\nu )},`$ (15b)
$`\overline{\mathrm{\Gamma }}_{\mu \nu }^\alpha \widehat{\mathrm{\Gamma }}_{\mu \nu }^\alpha `$ $`=\left({\displaystyle \frac{g}{1+2gX}}{\displaystyle \frac{b}{1+2bX}}\right)\psi ^\alpha _{(\mu }\psi _{\nu )}(g\overline{\mathrm{g}}^{\alpha \beta }b\widehat{\mathrm{g}}^{\alpha \beta })B_{\beta (\mu }\psi _{\nu )}.`$ (15c)
Taking the divergence of (13b) and using the Bianchi identity $`\overline{}_\nu \overline{G}^{\mu \nu }=0`$, results in
$$0=_\nu [\mu T^{\mu \nu }]+\mu T^{\beta \alpha }(\overline{\mathrm{\Gamma }}_{\alpha \beta }^\mu \mathrm{\Gamma }_{\alpha \beta }^\mu )+\kappa \widehat{\mu }\widehat{T}^{\beta \alpha }(\overline{\mathrm{\Gamma }}_{\alpha \beta }^\mu \widehat{\mathrm{\Gamma }}_{\alpha \beta }^\mu ),$$
(16)
where we have re-written the covariant derivative $`\overline{}_\nu `$ in terms of $`_\nu `$ when acting on the vector field energy-momentum tensor, and in terms of $`\widehat{}_\nu `$ when acting on the matter energy-momentum tensor. We have also used (3) to drop the divergence that results from the latter. The goal here is to show that this equation is automatically satisfied, i.e., it does not entail any additional restrictions on the matter fields or $`\psi _\mu `$ that are not already in existence in their respective field equations.
Taking the divergence of (8) and using $`_\alpha B_{\beta \gamma }+_\beta B_{\gamma \alpha }+_\gamma B_{\alpha \beta }=0`$ and (13a), we find
$$_\nu [\mu T^{\mu \nu }]=\mu \psi ^\mu _\nu [V^{}\psi ^\nu ]+(g\mu T^{\alpha \beta }+\kappa (gb)\widehat{\mu }\widehat{T}^{\alpha \beta })B_{}^{\mu }{}_{\alpha }{}^{}\psi _\beta .$$
(17)
Now taking the divergence of (13a) and using (17) and (15), we find that
$$\begin{array}{c}\mu (1+2gX)_\nu [V^{}\psi ^\nu ]+g\mu T^{\alpha \beta }(_\alpha \psi _\beta +g\psi _\alpha \psi ^\mu B_{\mu \beta })\hfill \\ \hfill +\kappa (gb)\widehat{\mu }\widehat{T}^{\alpha \beta }(\widehat{}_\alpha \psi _\beta +g\psi _\alpha \psi ^\mu B_{\mu \beta })=0.\end{array}$$
(18)
Using this to replace $`_\nu [V^{}\psi ^\nu ]`$ in (17) and identifying the various connections from (15), we find that the right-hand side of (16) vanishes identically. Thus we know that any matter model that conserves energy-momentum with respect to $`\widehat{\mathrm{g}}_{\mu \nu }`$ is consistent with the gravitational structure that we have introduced.
## 4. Causality and energy conditions
The โmost physicalโ metric is clearly $`\widehat{\mathrm{g}}_{\mu \nu }`$, since it describes the geometry on which matter propagates and interacts (c.f., the argument following (3) and ). Because all matter fields are coupled to the same metric $`\widehat{\mathrm{g}}_{\mu \nu }`$ in exactly the same way, the weak equivalence principle is satisfied. Furthermore, because one can work in a local Lorentz frame of $`\widehat{\mathrm{g}}_{\mu \nu }`$, in which non-gravitational physics takes on its special relativistic form, the Einstein equivalence principle is also satisfied. However, because $`\widehat{\mathrm{g}}_{\mu \nu }`$ does not couple to matter in the same way as in general relativity unless $`\psi _\mu =0`$, the strong equivalence principle will be violated. Clearly, we could introduce couplings of the form $`\psi _\mu \widehat{J}_{\mathrm{matter}}^\mu `$, where $`\widehat{J}_{\mathrm{matter}}^\mu `$ is a current constructed from the matter fields, into the action that would generically violate all equivalence principles.
In this type of model one cannot speak of causality without some qualification. From (11), $`\widehat{\mathrm{g}}_{\mu \nu }`$ and $`\overline{\mathrm{g}}_{\mu \nu }`$ are related by
$$\overline{\mathrm{g}}_{\mu \nu }=\widehat{\mathrm{g}}_{\mu \nu }+(gb)\psi _\mu \psi _\nu ,\overline{\mathrm{g}}^{\mu \nu }=\widehat{\mathrm{g}}^{\mu \nu }\frac{(gb)}{1+(gb)\widehat{\psi }^2}\widehat{\psi }^\mu \widehat{\psi }^\nu .$$
(19)
Characteristic surfaces of the field equations are determined from each metric by $`\overline{\mathrm{g}}^{\alpha \beta }V_\alpha V_\beta =0`$ and $`\widehat{\mathrm{g}}^{\alpha \beta }V_\alpha V_\beta =0`$, where $`V_\alpha `$ is a covector field that lies on the null cone of $`\overline{\mathrm{g}}^{\mu \nu }`$ and $`\widehat{\mathrm{g}}^{\mu \nu }`$, respectively. Because these two metrics are generally different and define different characteristic surfaces, we must specify, for example, a $`\overline{\mathrm{g}}`$-timelike vector $`V_\mu `$ as being one that satisfies $`\overline{\mathrm{g}}^{\alpha \beta }V_\alpha V_\beta >0`$. Because of this extended notion of causality, we will call a vector timelike if it is timelike with respect to all metrics in the theory, and spacelike if it is spacelike with respect to all the metrics in the theory .
Equation (19) allows us to connect these different causal relationships. Considering some covector $`V_\mu `$, we have
$$\overline{\mathrm{g}}^{\alpha \beta }V_\alpha V_\beta =\widehat{\mathrm{g}}^{\alpha \beta }V_\alpha V_\beta \frac{(gb)}{1+(gb)\widehat{\psi }^2}(\widehat{\mathrm{g}}^{\alpha \beta }V_\alpha \psi _\beta )^2,$$
(20)
and if we assume that $`V_\mu `$ is $`\widehat{\mathrm{g}}`$-null, then it is generically (assuming that $`\widehat{\mathrm{g}}(\psi ,V)0`$, i.e., $`\psi _\mu `$ and $`V_\mu `$ are not relatively $`\widehat{\mathrm{g}}`$-null)
$$\{\begin{array}{cc}\overline{\mathrm{g}}\text{timelike}\hfill & b>g,\hfill \\ \overline{\mathrm{g}}\text{null}\hfill & b=g,\hfill \\ \overline{\mathrm{g}}\text{spacelike}\hfill & b>g.\hfill \end{array}$$
(21)
The main motivation for considering these theories is that they should have something to say about the horizon problem in the early universe. If $`\psi _\mu 0`$, then if we choose $`b>g`$, matter fields will propagate outside the light cone of the gravitational field as illustrated in Figure 1.
As $`\psi _\mu 0`$ the matter light cone will โcontractโ and matter and gravitational disturbances will eventually propagate at the same velocity. If one considers a frame in which gravitational waves propagate at a constant speed, then as the light cone of matter contracts, the universe will appear to expand acausally to material observers. This is illustrated in Figure 2.
To get some feel for how this happens, and what other consequences there are, we write (13b) in terms of the Ricci curvature of $`\overline{\mathrm{g}}_{\mu \nu }`$:
$$\overline{R}^{\mu \nu }=\frac{1}{2}\frac{\mu }{\overline{\mu }}\left(T^{\mu \nu }\frac{1}{2}\overline{\mathrm{g}}^{\mu \nu }\overline{\mathrm{g}}_{\alpha \beta }T^{\alpha \beta }\right)+\frac{1}{2}\kappa \frac{\widehat{\mu }}{\overline{\mu }}\left(\widehat{T}^{\mu \nu }\frac{1}{2}\overline{\mathrm{g}}^{\mu \nu }\overline{\mathrm{g}}_{\alpha \beta }\widehat{T}^{\alpha \beta }\right).$$
(22)
The condition that the presence of matter causes convergence of $`\overline{\mathrm{g}}`$-timelike or $`\overline{\mathrm{g}}`$-null geodesic congruences is that $`\overline{R}_{\mu \nu }V^\mu V^\nu 0`$ for $`V^\mu `$ a $`\overline{\mathrm{g}}`$-timelike or $`\overline{\mathrm{g}}`$-null vector. If we assume that $`\widehat{\psi }^\mu `$ is a timelike vector (which is a reasonable cosmological assumption) and consider the special case $`V^\mu =\sigma \widehat{\psi }^\mu `$, we find the following result:
$$\begin{array}{cc}\hfill \overline{R}^{\mu \nu }\overline{V}_\mu \overline{V}_\nu & =\frac{1}{2}\frac{\mu }{\overline{\mu }}\sigma ^2\frac{1+g\psi ^2}{(1+b\psi ^2)^2}\left[๐ฏ_\psi \frac{1}{2}(b2g)\psi ^2T^{\alpha \beta }\psi _\alpha \psi _\beta \right]\hfill \\ & +\frac{1}{2}\kappa \frac{\widehat{\mu }}{\overline{\mu }}\sigma ^2\left(1+(gb)\widehat{\psi }^2\right)\left[\widehat{๐ฏ}_\psi \frac{1}{2}(bg)\widehat{\psi }^2\widehat{T}(\psi ,\psi )\right].\hfill \end{array}$$
(23)
where $`\widehat{T}:=\widehat{T}^{\mu \nu }\widehat{\mathrm{g}}_{\mu \nu }`$ and $`T:=T^{\mu \nu }\mathrm{g}_{\mu \nu }`$. In (23) we have written $`๐ฏ_\psi =T^{\alpha \beta }\psi _\alpha \psi _\beta \frac{1}{2}\psi ^2T`$ and $`\widehat{๐ฏ}_\psi =\widehat{T}^{\alpha \beta }\widehat{\psi }_\alpha \widehat{\psi }_\beta \frac{1}{2}\widehat{\psi }^2\widehat{T}`$, both of which are non-negative if the strong energy condition is satisfied by the vector field and matter (in the matter frame), respectively. If $`b>2g`$ then the second contribution from each energy-momentum tensor will be non-positive (assuming that each also satisfies the weak energy condition), and we find that a defocusing of geodesics is possible.
This implies that in the presence of a nonzero $`\psi _\mu `$, the singularity theorems as applied to $`\overline{\mathrm{g}}_{\mu \nu }`$ are weakened. As we shall see later, this will presumably have no effect on the collapse to a black hole in the present epoch, since in the vector field models the cosmological solution requires that $`\psi _\mu =0`$ when the matter energy density falls below a threshold. It is nevertheless reasonable to expect that the presence of $`\psi _\mu `$, and the light cone fluctuations that result, would have an effect on the threshold of black hole formation .
This result is also unfortunately not so clear as far as its physical implications. Although we can speak of singularity theorems as applied to $`\overline{\mathrm{g}}_{\mu \nu }`$, because the field equations that determine it are very similar to those of GR, it is the metric $`\widehat{\mathrm{g}}_{\mu \nu }`$ that is physically measured by the propagation of material test bodies. Indeed, the issue of how to identify a singularity is exacerbated in these models, because one can imagine the possibility that $`\widehat{\mathrm{g}}_{\mu \nu }`$ is well-behaved, whereas $`\overline{\mathrm{g}}_{\mu \nu }`$ and $`\psi _\mu `$ individually are not. The main point is that โordinaryโ matter can contribute to the field equations for $`\overline{\mathrm{g}}_{\mu \nu }`$ as if it violates the strong energy condition. We will see exactly this happening in the following section.
## 5. Cosmology
Implicit in the idea of a varying lightspeed is that the speed of light is changing with respect to some fixed frame of reference. If one introduces a fundamental frame for this, then it is perhaps sensible to introduce a function $`c(t,x)`$ to describe this variability . The models that we are considering are based on the idea that the speed of light can be changing with respect to the speed of gravitational disturbances, and therefore any indication of the speed of light as a function of spacetime is frame-dependent. In particular, we will see that a frame in which the speed of light is constant and the speed of gravitational disturbances is changing is connected via a diffeomorphism to a frame where the speed of gravitational disturbances is constant, and the speed of light is changing (c.f., Figure 2). We will derive the quantities of interest (the local light cone, horizons, etc.) directly from the relevant metric, thereby avoiding any guesswork as to which โspeed of lightโ to useโthe gravitational or electromagnetic . The constant $`c`$ is fixed in the present universe by making measurements of the electromagnetic field.
In a homogeneous and isotropic (FRW) universe, the vector field $`\psi _\mu `$ has components $`\psi _\mu =(c\psi _0(\tau ),0,0,0)`$. We will begin with the metric $`\mathrm{g}_{\mu \nu }`$ in comoving form
$$\mathrm{g}_{\mu \nu }dx^\mu dx^\nu =c^2d\tau d\tau R(\tau )^2\gamma _{ij}dx^idx^j,$$
(24a)
and therefore
$`\widehat{\mathrm{g}}_{\mu \nu }dx^\mu dx^\nu `$ $`=\widehat{\mathrm{\Theta }}^2c^2d\tau d\tau R(\tau )^2\gamma _{ij}dx^idx^j,`$ (24b)
$`\overline{\mathrm{g}}_{\mu \nu }dx^\mu dx^\nu `$ $`=\overline{\mathrm{\Theta }}^2c^2d\tau d\tau R(\tau )^2\gamma _{ij}dx^idx^j.`$ (24c)
The spatial metric in spherical coordinates has the standard form
$$\gamma _{ij}=\mathrm{diag}(1/(1kr^2),r^2,r^2\mathrm{sin}^2\theta ),$$
(25)
and we have defined
$$\widehat{\mathrm{\Theta }}:=\sqrt{1+2bX},\text{and}\overline{\mathrm{\Theta }}:=\sqrt{1+2gX},$$
(26)
where from (7) we have $`X=\frac{1}{2}\psi _0^2`$.
Although we begin with the choice (24a), once we have derived the field equations (which is significantly easier to do in this gauge) we will make a coordinate transformation in order to put $`\widehat{\mathrm{g}}_{\mu \nu }`$ in comoving form and thereby make a comparison with the standard cosmological results a simpler matter. Note that we are reversing the definitions of $`t`$ and $`\tau `$ as used in our previous article .
The matter energy-momentum tensor will have a perfect fluid form:
$$\widehat{T}^{\mu \nu }=\left(\rho +\frac{p}{c^2}\right)\widehat{u}^\mu \widehat{u}^\nu p\widehat{\mathrm{g}}^{\mu \nu },$$
(27)
where we have written the velocity field as $`\widehat{u}^\mu `$ to emphasize that it is normalized using the metric $`\widehat{\mathrm{g}}_{\mu \nu }`$, so that
$$\widehat{\mathrm{g}}_{\mu \nu }\widehat{u}^\mu \widehat{u}^\nu =c^2.$$
(28)
This, we believe, is the origin of the confusion in , in which we have been accused of making an algebraic mistake where there is none. In that work, the authors appear to have normalized the vector field as $`\mathrm{g}_{\mu \nu }\widehat{u}^\mu \widehat{u}^\nu =c^2`$. The resulting form for $`\widehat{T}^{\mu \nu }`$ clearly will not satisfy (3), and as we have seen in Section 3, will lead to inconsistent field equations since the Bianchi identities will now lead to nontrivial constraints on the matter fields.
The correct choice (28) leads to $`\widehat{u}^0=1/\widehat{\mathrm{\Theta }}`$, and therefore
$$\widehat{T}^{00}=\frac{\rho }{\widehat{\mathrm{\Theta }}^2},\widehat{T}^{0i}=0,\widehat{T}^{ij}=\frac{p}{R^2}\gamma ^{ij}.$$
(29)
Thus equation $`(58)`$ of is incorrect; the first term on the right-hand side should be $`\rho /\sqrt{1+\beta \psi _0^2}`$ rather than $`\sqrt{1+\beta \psi _0^2}\rho `$.
The matter conservation laws (3) lead to the usual relation
$$_\tau \rho +3\frac{_\tau R}{R}\left(\rho +\frac{p}{c^2}\right)=0.$$
(30)
Note that this form also holds with any choice of time coordinate $`\tau \tau (t)`$.
The components of the energy-momentum tensor for $`\psi `$ are
$$T^{00}=\frac{1}{c^2}(2XV^{}V),T^{0i}=0,T^{ij}=V\frac{1}{R^2}\gamma ^{ij},$$
(31)
and using the components of the Einstein tensor
$`\overline{G}^{00}`$ $`={\displaystyle \frac{3}{c^4\overline{\mathrm{\Theta }}^4}}\left[\left({\displaystyle \frac{_\tau R}{R}}\right)^2+{\displaystyle \frac{kc^2\overline{\mathrm{\Theta }}^2}{R^2}}\right],`$ (32a)
$`\overline{G}^{ij}`$ $`=\gamma ^{ij}{\displaystyle \frac{1}{c^2R^2\overline{\mathrm{\Theta }}^2}}\left[2{\displaystyle \frac{_\tau ^2R}{R}}+\left({\displaystyle \frac{_\tau R}{R}}\right)^2+{\displaystyle \frac{kc^2\overline{\mathrm{\Theta }}^2}{R^2}}2{\displaystyle \frac{_\tau R}{R}}{\displaystyle \frac{_\tau \overline{\mathrm{\Theta }}}{\overline{\mathrm{\Theta }}}}\right],`$ (32b)
from (13b) we have the Friedmann equations
$$\left(\frac{_\tau R}{R}\right)^2+\frac{kc^2\overline{\mathrm{\Theta }}^2}{R^2}=\frac{\kappa c^4}{6}\overline{\mathrm{\Theta }}^3\left[\frac{\rho }{\widehat{\mathrm{\Theta }}}+\frac{1}{\kappa c^2}(2XV^{}V)\right],$$
(33a)
$$2\frac{_\tau ^2R}{R}+\left(\frac{_\tau R}{R}\right)^2+\frac{kc^2\overline{\mathrm{\Theta }}^2}{R^2}2\frac{_\tau R}{R}\frac{_\tau \overline{\mathrm{\Theta }}}{\overline{\mathrm{\Theta }}}=\frac{\kappa c^2}{2}\overline{\mathrm{\Theta }}\left[\widehat{\mathrm{\Theta }}p+\frac{1}{\kappa }V\right].$$
(33b)
The single remaining Proca field equation from (13a) is
$$\frac{1}{c\widehat{\mathrm{\Theta }}}\psi _0\left[\widehat{\mathrm{\Theta }}\left(\overline{\mathrm{\Theta }}^2V^{}gV\right)\kappa (bg)c^2\rho \right]=0.$$
(33c)
We now perform the coordinate transformation
$$dt=\widehat{\mathrm{\Theta }}d\tau ,$$
(34)
and defining
$$\eta =\frac{\overline{\mathrm{\Theta }}}{\widehat{\mathrm{\Theta }}}=\sqrt{\frac{1+2gX}{1+2bX}},$$
(35)
we see that the metric $`\widehat{\mathrm{g}}_{\mu \nu }`$ is put into comoving form
$`\widehat{\mathrm{g}}_{\mu \nu }dx^\mu dx^\nu `$ $`=c^2dtdtR^2(t)\gamma _{ij}dx^idx^j,`$ (36a)
$`\overline{\mathrm{g}}_{\mu \nu }dx^\mu dx^\nu `$ $`=\eta ^2(t)c^2dtdtR^2(t)\gamma _{ij}dx^idx^j.`$ (36b)
Making the change of time coordinate directly on (33), (33c) is unchanged, and we find
$`\left({\displaystyle \frac{\dot{R}}{R}}\right)^2+{\displaystyle \frac{kc^2\eta ^2}{R^2}}`$ $`=\eta ^2{\displaystyle \frac{\kappa c^4}{6}}\rho _{\mathrm{eff}},`$ (37a)
$`2{\displaystyle \frac{\ddot{R}}{R}}+\left({\displaystyle \frac{\dot{R}}{R}}\right)^2+{\displaystyle \frac{kc^2\eta ^2}{R^2}}2{\displaystyle \frac{\dot{R}}{R}}{\displaystyle \frac{\dot{\eta }}{\eta }}`$ $`=\eta ^2{\displaystyle \frac{\kappa c^2}{2}}p_{\mathrm{eff}},`$ (37b)
where we will write, for example $`\dot{\rho }=_t\rho `$. In (37), we have defined the effective energy and pressure densities as
$$\rho _{\mathrm{eff}}=\eta \left(\rho +\frac{1}{\kappa c^2}\widehat{\mathrm{\Theta }}(2XV^{}V)\right),p_{\mathrm{eff}}=\frac{1}{\eta }\left(p+\frac{1}{\kappa \widehat{\mathrm{\Theta }}}V\right).$$
(38)
The reason for making these definitions is that (37a) has exactly the form of the Friedmann equations for the metric $`\overline{\mathrm{g}}_{\mu \nu }`$ in (36b), and therefore these effective energy and momentum densities will also satisfy the conservation laws (30).
Some careful observations are in order. The function $`R(t)`$ is determined from (37a) in which an (unusual) effective energy density appears. It also formally appears to be a โvarying constantsโ model with $`c(t)=\eta (t)c`$ and $`G(t)=\eta ^2(t)G`$ coupled to $`\rho _{\mathrm{eff}}`$. We could equally well expand $`\rho _{\mathrm{eff}}`$ to find a varying constants form with
$$c(t)=\eta c,G(t)=\eta ^3G,\mathrm{\Lambda }(t)=\frac{1}{2}\overline{\mathrm{\Theta }}(2XV^{}V),$$
(39)
or, if we interpret (37a) as a varying constants theory written in a non-comoving coordinate system with metric of the form (36b), we would have
$$c(t)=c,G(t)=\eta G,\mathrm{\Lambda }(t)=\frac{1}{2}\overline{\mathrm{\Theta }}(2XV^{}V),$$
(40)
and, as we will see in Section 5.2, once the vector field equation (33c) is solved, the resulting system need not take on an FRW form at all. Clearly any such identification is ambiguous.
Note though, that the resulting function $`R(t)`$ appears in (36a), which is written in comoving coordinates, and therefore the speed of light is constant. This emphasizes that having a โvarying speed of lightโ is a frame-dependent statement. In a frame where the speed of matter propagation (including electromagnetic fields) is constant, the speed of gravitational waves will be changing. In a frame where the speed of gravitational waves is constant, the speed of matter propagation will be changing. This, of course, is as it should be, since we have not introduced any nondynamical preferred frame into our model.
In the following we will specialize to a model where the vector field potential is a simple mass term:
$$V=m^2X,V^{}=m^2.$$
(41)
In this case (38) becomes
$$\rho _{\mathrm{eff}}=\eta \left(\rho +(bg)\rho _{\mathrm{๐๐ก}}\widehat{\mathrm{\Theta }}X\right),p_{\mathrm{eff}}=\frac{1}{\eta }\left(p+c^2(bg)\rho _{\mathrm{๐๐ก}}\frac{X}{\widehat{\mathrm{\Theta }}}\right),$$
(42)
and we find for later use that
$$\rho _{\mathrm{eff}}+\frac{3}{c^2}p_{\mathrm{eff}}=\eta \rho +\frac{3}{c^2\eta }p+(bg)\rho _{\mathrm{๐๐ก}}\left(\overline{\mathrm{\Theta }}X+\frac{3}{X\overline{\mathrm{\Theta }}}\right).$$
(43)
The nontrivial solution ($`\psi _00`$) of the field equation (33c) leads to
$$\rho =\rho _{\mathrm{๐๐ก}}\widehat{\mathrm{\Theta }}(1+gX),$$
(44)
where
$$\rho _{\mathrm{๐๐ก}}=\frac{m^2}{\kappa c^2(bg)},H_{\mathrm{๐๐ก}}=\sqrt{\frac{c^2m^2}{6(bg)}},$$
(45)
are the density at which $`\psi _0^2=0`$ is reached, and the inverse Hubble time at which this occurs (assuming that $`k=0`$).
We can now write the acceleration parameter as observed by material observers from (37) as
$$\widehat{q}=\frac{\ddot{R}}{H^2R}=\frac{\kappa c^4}{12}\frac{\eta ^2}{H^2}\left(\rho _{\mathrm{eff}}+\frac{3}{c^2}p_{\mathrm{eff}}\right)\frac{\dot{\eta }}{H\eta },$$
(46)
where we have defined the Hubble function $`H=\dot{R}/R`$. Taking a derivative of the definition (35), relating the result to $`\dot{\rho }`$ using the derivative of (44), and removing $`\dot{\rho }`$ using (30), we find
$$\frac{\dot{\eta }}{H\eta }=3\frac{(bg)}{\rho _{\mathrm{๐๐ก}}\overline{\mathrm{\Theta }}^2\widehat{\mathrm{\Theta }}(g+b+3bgX)}\left(\rho +\frac{p}{c^2}\right).$$
(47)
We will return to this shortly.
### 5.1. The very early universe
For very short times following the initial singularity, we expect that $`\psi _0`$ is large, and if we assume that $`gX1`$ and $`bX1`$, then from (44) we find that
$$\rho =\rho _{\mathrm{๐๐ก}}\sqrt{2b}gX^{3/2}.$$
(48)
This results in the Friedmann equation
$$\left(\frac{\dot{R}}{R}\right)^2+\frac{k\overline{c}^2}{R^2}=\frac{\overline{\kappa }\overline{c}^4}{6}\rho ,$$
(49)
where
$$\overline{c}=c\sqrt{\frac{g}{b}},\overline{G}=G\sqrt{\frac{g}{b}},\overline{\kappa }=\frac{16\pi \overline{G}}{\overline{c}^4}.$$
(50)
Although the behaviour of the solutions to (49) are well-known , it is worth pointing out that the โeffectiveโ constants $`\overline{c}`$ and $`\overline{G}`$ are not interpretable as the effective speed of light and gravitational constant, rather they are effective constants that dictate how the gravitational field reacts to the presence of matter. Matter fields continue to propagate with speed $`c`$ consistent with (36a). It is the gravitational field perturbations that propagate with speed $`\overline{c}`$, which is the justification for the notation.
During this phase there is clearly no inflation, but the horizon scales of the gravitational field and matter fields are related by:
$$\overline{d}_H(t)=\frac{\overline{c}}{c}\widehat{d}_H(t)=\sqrt{\frac{g}{b}}\widehat{d}_H(t),\text{where}\widehat{d}_H(t)=cR(t)_0^t\frac{ds}{R(s)},$$
(51)
with a similar definition for $`\overline{d}_H(t)`$ using the metric (36b). Because we have $`g<b`$ we expect that not only is the speed of gravitational disturbances slower than that of matter, but also that the coupling between matter and the gravitational sector is also lessened.
What we have here is very close to what was originally envisaged by one of us in . This is part of the motivation for including the $`g0`$ possibility, the other is that the approach to the initial singularity in this phase follows the same path as in ordinary GR+matter models, with a re-interpretation of the parameters. In this case we have a model that interpolates between this initial period where $`\overline{c}>c`$ and the later universe where $`\overline{c}=c`$. We turn to this next.
### 5.2. Inflation and light cone contraction
As $`\psi _0`$ decreases towards the point where $`gX1`$ the solution found in Section 5.1 will no longer be a good approximation. If we now consider the solution when $`gX1`$, from (44) we have
$$\widehat{\mathrm{\Theta }}=\frac{\rho }{\rho _{\mathrm{๐๐ก}}},\text{or}X=\frac{1}{2b}\left[\left(\frac{\rho }{\rho _{\mathrm{๐๐ก}}}\right)^21\right],$$
(52)
and the Friedmann equation (37a) becomes
$$\left(\frac{\dot{R}}{R}\right)^2+\frac{kc^2\eta ^2}{R^2}\left(\frac{\rho _{\mathrm{๐๐ก}}}{\rho }\right)^2=\frac{\kappa c^4}{12}\rho _{\mathrm{๐๐ก}}\left[1+\left(\frac{\rho _{\mathrm{๐๐ก}}}{\rho }\right)^2\right].$$
(53)
In this limit
$$\rho _{\mathrm{eff}}+\frac{3}{c^2}p_{\mathrm{eff}}=\frac{1}{\rho _{\mathrm{๐๐ก}}}\left[\rho \left(\rho +\frac{3}{c^2}p\right)+\rho ^2\rho _{\mathrm{๐๐ก}}^2\right],$$
(54)
which is greater than zero if the strong energy condition is satisfied, since $`\rho \rho _{\mathrm{๐๐ก}}`$, and (46) reduces to
$$\widehat{q}=\frac{\kappa c^4\rho _{\mathrm{๐๐ก}}}{12H^2}\left[\frac{1}{\rho }\left(\rho +\frac{3}{c^2}p\right)+1\left(\frac{\rho _{\mathrm{๐๐ก}}}{\rho }\right)^2\right]\frac{3}{\rho }\left(\rho +\frac{p}{c^2}\right).$$
(55)
Since we expect that $`H^2`$ is large in the early universe (we can arrange that $`\rho _{\mathrm{๐๐ก}}\rho _c`$ where $`\rho _c=12H^2/(\kappa c^4)`$), it is clear from (55) that even if matter satisfies the strong energy condition, the final term will dominate and $`\widehat{q}<0`$ (unless, perhaps, the weak energy condition is also violated). This is the expansion of the universe as seen by material observers. The acceleration of the gravitational geometry $`\overline{q}`$ would lack the final term and therefore $`\overline{q}>0`$.
That we get inflation was demonstrated previously , where an exact solution for $`k=0`$ and $`g=0`$ was found. Although we discovered that we could not get enough expansion to solve the horizon problem with pure radiation, a slowly rolling scalar field could provide the necessary negative pressure. The role that the extra structure of our model plays is that the fine-tuning that is required in a simple scalar field, potential-driven model is alleviated.
The flatness problem requires a bit more explanation. Dividing (37a) by $`H^2`$ and defining
$$\overline{ฯต}=\frac{kc^2\eta ^2}{(\dot{R})^2},$$
(56)
we find a differential equation that $`\overline{ฯต}`$ satisfies by taking a derivative and using (37) to give
$$\dot{\overline{ฯต}}=\frac{\kappa c^4\eta ^2}{6H}\overline{ฯต}\left(\rho _{\mathrm{eff}}+\frac{3}{c^2}p_{\mathrm{eff}}\right).$$
(57)
Therefore, since $`\overline{ฯต}>0`$ and $`H>0`$ in the early universe, the only way for $`\overline{ฯต}=0`$ to be an attractor for (37a) is for $`\rho _{\mathrm{eff}}+\frac{3}{c^2}p_{\mathrm{eff}}<0`$ at least for part of the history of the universe. What is not so obvious is whether the quantity $`\overline{ฯต}`$ as defined in (56) is of physical relevance.
Following the quantity of geometrical importance is the $`3`$-curvature of the spacelike slices: $`6k/R^2`$, which suggests that the physically meaningful quantity to examine would be
$$\widehat{ฯต}=\frac{kc^2}{(\dot{R})^2},$$
(58)
which has the equation of motion
$$\dot{\widehat{ฯต}}=2\widehat{ฯต}\widehat{q}.$$
(59)
Another way of stating this is that the curvature radius defines $`\mathrm{\Omega }`$ through :
$$R_{\mathrm{๐๐ข๐๐ฃ}}=\frac{R}{|k|^{1/2}}=\frac{c}{H(|\mathrm{\Omega }1|)^{1/2}},$$
(60)
and so $`\widehat{ฯต}=|\mathrm{\Omega }1|`$. Since we found from (55) that $`\widehat{q}<0`$ in the early universe, clearly $`\widehat{ฯต}=0`$ is an attractor for (37a), and since it is most-likely the quantity of physical importance for matter physics, we can also claim to have solved the flatness problem once the horizon problem is solved.
Some comments on the remarks made in on the flatness problem are in order. If we look at the acceleration of the universe from the point of view of gravitational โobserversโ only, then the universe will not appear to inflate and the flatness problem is not solved. This appears to be the point of view advocated implicitly in . The point we are making here is that the geometry that we observe as material bodies is that described by $`\widehat{\mathrm{g}}_{\mu \nu }`$, and as far as material observers are concerned the universe will appear to inflate. As it does so, the spatial curvature that would be measured by such observers necessarily increases. That is, the curvature radius sets the length scale of the size of spatial separations on which the effects of spatial curvature become important. That, of course, is the rationale behind the statement that โthe horizon and flatness problems are geometrically linkedโ.
## 6. conclusions
One of our purposes here was to give a more complete description of the model than was given in . In doing so we showed that the universe generically accelerates ($`\widehat{q}<0`$) during some period in the early universe, and that in the same period the physical importance of spatial curvature diminishes ($`|\mathrm{\Omega }1|`$ is decreasing). This can occur even when the matter fields satisfy the strong energy condition. This conclusion is the opposite of that which appeared recently , who take a somewhat different point of view on the interpretation of the metrics appearing in these theories. Nevertheless, we have demonstrated conclusively that the claim appearing in that we had made an algebraic error in is false.
The model that we have considered generalizes that which appeared in in a way that more closely follows the scenario discussed in . In the very early universe, matter and gravitational fields propagate with different and approximately constant velocities. During a period during which the matter light cone, originally much larger than the light cone of gravity, contracts, material observers will see an acausal expansion of the universe similar to inflation. Because the light cone of gravity does not undergo the same contraction, we expect there to be an observable difference in the scalar versus tensor contributions to the cosmic microwave background anisotropies.
## Acknowledgements
J.W.M. is supported by the Natural Sciences and Engineering Research Council of Canada. |
warning/0003/physics0003097.html | ar5iv | text | # VECTOR CONSTANTS OF MOTION FOR TIME-DEPENDENT KEPLER AND ISOTROPIC HARMONIC OSCILLATOR POTENTIALS
## 1 Introduction
It is well known that in classical mechanics the knowledge of all first integrals of motion of a given problem is equivalent to finding its complete solution. Nowadays the search for first integrals has assumed an increasing importance in the determination of the integrability of a dynamical system. It is extremely important to know if a non-linear dynamical system will present chaotic behavior in some regions of the phase space. The notion of integrability is related to the existence of first integrals of motion. Several methods of finding first integrals are available in the literature for example, Lieโs method , Noetherโs theorem , or the direct method . Even if not all first integrals of motion associated with the problem at hand are found, it may happen that the ones which are obtained contribute to the discovery of the solution we are seeking for. Nevertheless, if we do find the solution we are after by solving the equations of motion in a straightforward way, it still may be profitable to look for additional constants of motion. Such is the case of the Kepler problem where the knowledge of the Laplace-Runge-Lenz vector , allows us to obtain the orbit in a simple way.
Of the inexhaustible wealth of problems which we can find in classical mechanics one of the most aesthetically appealing and important is the central field problem. Energy and angular momentum associated with this type of field are well known conserved quantities. However, other vector and tensor conserved quantities have been associated with some particular central fields. The Laplace-Runge-Lenz vector is a vector first integral of motion for the Kepler problem; the Fradkin tensor is conserved for the case of the harmonic oscillator and for any central field it is possible to find a vector first integral of motion as was shown in . In the general case these additional integrals of motion turn out to be complicated functions of the position $`๐ซ`$ and linear momentum $`๐ฉ`$ of the particle probing the central field. When orbits are closed and degenerated with respect to the mechanical energy, however, we should expect these additional constant of motion to be simple function of $`๐ซ`$ and $`๐ฉ`$. In this article we wish to exploit further this line of reasoning by determining the existence of such additional vector first integrals of motion for the time-dependent Kepler and isotropic harmonic oscillator problems. In particular, we will show that for the time-dependent Kepler problem the existence of a vector constant of motion coupled to a simple transformation of variables turns the problem easily integrable.
The structure of this paper goes as follows: in section 2 we establish the conditions which guarantee the existence of a vector first integral for a general central force field. In section 3 we put the method to test by rederiving some well known results such as the conservation of angular momentum in an arbitrary central field, the conservation of the Laplace-Runge-Lenz vector for the Kepler problem, and the conservation of the Fradkin tensor fixing en route these specific fields to which they correspond. In section 4 we consider the time-dependent case establishing generalizations of the examples considered before and presenting new ones. In section 5 we show that the existence of a vector first integral enable us to find the orbits of a test particle. This is accomplished for the case of harmonic oscillator, and the time-dependent Kepler problem. Also the period of the time-dependent Kepler problem is obtained. Finally, section 6 is reserved for final comments.
## 2 Constructing vector constants of motion
The force $`๐(๐ซ,t)`$ acting on a test particle moving in a central but otherwise arbitrary and possible time-dependent field of force $`g(r,t)`$ can be written as
$$๐(๐ซ,t)=g(r,t)๐ซ,$$
(1)
where $`๐ซ=๐ซ\left(t\right)`$ is the position vector with respect to the center of force, $`r`$ is its magnitude, and $`t`$ is the time. To this test particle we assume that it is possible to associate a vector $`๐ฃ`$ which in principle can be written in the form
$$๐ฃ(๐ฉ,๐ซ,t)=A(๐ฉ,๐ซ,t)๐ฉ+B(๐ฉ,๐ซ,t)๐ซ,$$
(2)
where $`๐ฉ=๐ฉ\left(t\right):=m\dot{๐ซ}`$ $`\left(t\right)`$ is the linear momentum, $`m`$ is the reduced mass and $`A`$, $`B`$ are arbitrary scalar functions of $`๐ฉ,๐ซ`$ and $`t`$. Taking the total time derivative of (2) and making use of Newtonโs second law of motion we readily obtain
$$\frac{d๐ฃ}{dt}=\left(Ag+\frac{dB}{dt}\right)๐ซ+\left(\frac{dA}{dt}+\frac{B}{m}\right)๐ฉ.$$
(3)
If we assume that $`๐ฃ`$ is a constant of motion it follows that the functions $`A`$ and $`B`$ must satisfy
$$Ag+\frac{dB}{dt}=0,$$
(4)
$$\frac{dA}{dt}+\frac{B}{m}=0.$$
(5)
Eliminating $`B`$ between (4) and (5) we obtain
$$m\frac{d^2A}{dt^2}gA=0.$$
(6)
It follows from (5) that $`๐ฃ`$ can be written in the form
$$๐ฃ=A๐ฉm\frac{dA}{dt}๐ซ.$$
(7)
Therefore, since (6) is equivalent to both (4) and (5) if the field $`g(r,t)`$ is known any solution of (6) will yield a vector constant of motion of the form given by (7). Equation (6), however, is a differential equation whose solution may turn out to be a hard task to accomplish. Nevertheless, we can make progress if instead of trying to tackle it directly we make plausible guesses concerning $`A`$ thereby linking $`๐ฃ`$ to specific forms of the field $`g(r,t)`$. This procedure is tantamount to answering the following question: given $`๐ฃ`$ what type of central field will admit it as a constant of motion? The answer is given in the next section.
## 3 Simple examples
With $`๐ซ`$, $`๐ฉ`$ and a unit constant vector $`\widehat{๐ฎ}`$ we can construct the following scalars: $`\widehat{๐ฎ}๐ซ,`$ $`\widehat{๐ฎ}๐ฉ`$ and $`๐ซ๐ฉ.`$ Other possibilities will be considered later on. For the moment let us consider some simple possibilities for the scalar function $`A(๐ฉ,๐ซ,t)`$.
Consider first $`A(๐ฉ,๐ซ,t)=\widehat{๐ฎ}๐ซ.`$ It is immediately seen that this choice for $`A`$ satisfies (6) for any function $`g(r,t).`$ The constant vector $`๐ฃ`$ reads
$$๐ฃ=(\widehat{๐ฎ}๐ซ)๐ฉ(\widehat{๐ฎ}๐ฉ)๐ซ,$$
(8)
and it can be related to the angular momentum $`๐=๐ซ\times ๐ฉ`$ as follows. Firstly we recast (8) into the form
$$๐ฃ=๐\widehat{๐ฎ},$$
(9)
where $`๐=๐ฉ๐ซ๐ซ๐ฉ.`$ Since $`\widehat{๐ฎ}`$ is a constant vector we conclude that the constancy of $`๐ฃ`$ is equivalent to the constancy of $`๐`$ whose components are $`M_{jk}=p_jx_kx_jp_k`$, with $`i,j,k=1,2,3`$. The antisymmetrical tensor $`๐`$ is closely related to angular momentum $`๐`$ of the test particle. In fact, it can be easily shown that $`2L_i=\epsilon _{ijk}M_{jk}`$, where $`L_i`$ is the i-th angular momentum component and $`\epsilon _{ijk}`$ is the usual permutation symbol or Levi-Civita density. Therefore, this simple choice for $`A`$ leads to conservation of angular momentum for motion under a central arbitrary field $`g(r,t)`$.
Consider now the choice $`A(๐ฉ,๐ซ,t)=\widehat{๐ฎ}๐ฉ`$. Then making use of Newtonโs second law it follows that (7) is satisfied if we find a solution to
$$\frac{dg(r,t)}{dt}\widehat{๐ฎ}๐ซ=0.$$
(10)
For arbitrary values of $`\widehat{๐ฎ}๐ซ`$ we can find a solution to (10) if and only if $`\stackrel{.}{g(r,t)}0,`$ or $`g=g_0`$ where $`g_0`$ is a constant. In this case we can write
$$๐ฃ=(\widehat{๐ฎ}๐ฉ)๐ฉg_0(\widehat{๐ฎ}๐ซ)๐ซ.$$
(11)
If we choose the constant to be equal to $`k`$ then the central force field will correspond to an isotropic harmonic oscillator, $`๐=k๐ซ.`$ As before, (11) can be recasted into the form
$$๐ฃ=2m๐
\widehat{๐ฎ},$$
(12)
where $`๐
`$ is given by
$$๐
=\frac{๐ฉ๐ฉ}{2m}+k\frac{๐ซ๐ซ}{2}.$$
(13)
The tensor $`๐
`$ is symmetrical and is known as the Fradkin tensor . Finally, consider $`A=๐ซ๐ฉ.`$ For this choice of $`A`$ (7) yields
$$\frac{1}{g}\frac{dg}{dt}+\frac{3}{r}\frac{dr}{dt}=0,$$
(14)
where we have made use of (1) and also of the fact that $`d\widehat{๐ซ}/dt`$ and $`๐ซ`$ are perpendicular vectors. Equation (14) can be easily integrated if $`g`$ is considered to be a function of the radial distance $`r`$ only. If this is the case we obtain the Kepler field $`g(r)=k/r^3.`$ The constant vector $`๐ฃ`$ is then given by
$$๐ฃ=(๐ซ๐ฉ)๐ฉ(๐ฉ๐ฉ)๐ซ+mk\frac{๐ซ}{r},$$
(15)
Making use of a well known vector identity we can recast (15) into the form,
$$๐ฃ=๐\times ๐ฉmk\frac{๐ซ}{r}.$$
(16)
Therefore $`๐ฃ`$ can be equaled to minus the Laplace-Runge-Lenz vector $`๐`$. From (16) and the condition $`๐ฃ๐ซ=๐๐ซ=0`$ the allowed orbits for the Kepler problem can be obtained in a simple way, see for example .
## 4 Time-dependent fields
Let us now consider time-dependent central force fields for which we can build more general vector first integrals of motion. As with the time-independent case there are of course several possibilities when it comes to the choice of a function $`A`$ for a time-dependent central field. Here is one
$$A=\varphi (t)๐ซ๐ฉ+\psi (t)๐ซ๐ซ.$$
(17)
Evaluating the second derivative of (17) we obtain
$$\frac{d^2A}{dt^2}=\left(\frac{d^2\varphi }{dt^2}+\frac{4g\varphi }{m}+\frac{4}{m}\frac{d\psi }{dt}\right)๐ซ๐ฉ+\left(2\frac{d\varphi }{dt}g+\varphi \frac{dg}{dt}+\frac{d^2\psi }{dt^2}+\frac{2\psi g}{m}\right)๐ซ๐ซ$$
$$+\frac{2}{m^2}\left(m\frac{d\varphi }{dt}+\psi \right)๐ฉ๐ฉ.$$
(18)
Where we have made use of (1). If we impose the additional condition
$$\psi +m\frac{d\varphi }{dt}=0,$$
(19)
we eliminate the quadratic term in $`๐ฉ`$. With the condition given by (19) we can substitute for $`\psi `$ in (17) and (18) and take the results into (6) thus obtaining
$$3\left(m\frac{d^2\varphi }{dt^2}+g\varphi \right)๐ซ\frac{d๐ซ}{dt}+\left(\varphi \frac{dg}{dt}m\frac{d^3\varphi }{dt^3}+g\frac{d\varphi }{dt}\right)๐ซ๐ซ=0.$$
(20)
Equation (20) can be rewritten as
$$\frac{3}{2}\left(m\frac{d^2\varphi }{dt^2}+g\varphi \right)\frac{d(๐ซ^2)}{dt}+\left[\frac{d}{dt}\left(m\frac{d^2\varphi }{dt^2}+g\varphi \right)\right]๐ซ^2=0,$$
(21)
and easily integrated so as to yield
$$g(r,t)=\frac{m}{\varphi }\frac{d^2\varphi }{dt^2}+\frac{C}{\varphi r^3}.$$
(22)
The vector first integral of motion associated with (22) is
$$๐ฃ=\left(\varphi ๐ซ๐ฉm\frac{d\varphi }{dt}๐ซ๐ซ\right)๐ฉ+\left(m\frac{d\varphi }{dt}๐ซ๐ฉ\varphi ๐ฉ๐ฉ\frac{mC}{r}\right)๐ซ,$$
(23)
which can be simplified and written in the form
$$๐ฃ=m\varphi ^2๐\times \frac{d}{dt}\left(\frac{๐ซ}{\varphi }\right)\frac{mC๐ซ}{r}.$$
(24)
where we have made used of the fact the angular momentum is constant for any arbitrary central field whether it is time-independent or not. If in (24) we set $`\varphi =1`$ and $`C0`$, then from (22) we see that $`g(r,t)`$ is the Kepler field and $`๐ฃ`$ is minus the Laplace-Runge-Lenz vector as before; the scalar function $`A(๐ซ,๐ฉ,t)`$ reduces to $`๐ซ๐ฉ`$ which we have already employed in section 3. If we set $`(m/\varphi )d^2\varphi /dt^2=k(t),`$ that is, if $`\varphi `$ is an arbitrary function of the time, and also $`C=0`$, we have the time-dependent isotropic harmonic oscillator field, $`๐
\left(r\right)=k(t)๐ซ`$. In this case $`๐ฃ`$ is equal to the first term on the R.H.S. of (24). If $`(m/\varphi )d^2\varphi /dt^2=k`$ and $`C=0`$, we have the time-independent isotopic harmonic oscillator field but this time $`๐ฃ`$ is not the same vector as the one we have obtained before. The reason for this is our choice (17) for the scalar function $`A(๐ซ,๐ฉ,t)`$ which is not reducible to the form $`\widehat{๐ฎ}๐ฉ`$ employed previously.
As a last example let us consider again the time-dependent isotropic harmonic oscillator and show how it is possible to generalize the Fradkin tensor for this case. Let the function $`A(๐ซ,๐ฉ,t)`$ be written as
$$A=\varphi (t)\widehat{๐ฎ}๐ซ+\psi (t)\widehat{๐ฎ}๐ฉ.$$
(25)
The first and the second derivative of $`A`$ read
$$\frac{dA}{dt}=\frac{d\varphi }{dt}\widehat{๐ฎ}๐ซ+\frac{\varphi }{m}\widehat{๐ฎ}๐ฉ+\frac{d\psi }{dt}\widehat{๐ฎ}๐ฉ+g\psi \widehat{๐ฎ}๐ซ,$$
(26)
and
$$\frac{d^2A}{dt^2}=\left(\frac{d^2\varphi }{dt^2}+2g\frac{d\psi }{dt}+\frac{dg}{dt}\psi +\frac{g\varphi }{m}\right)\widehat{๐ฎ}๐ซ+\left(\frac{d^2\psi }{dt^2}+2g\frac{d\varphi }{dt}+\frac{g\psi }{m}\right)\widehat{๐ฎ}๐ฉ$$
(27)
Taking (27) into (6) we obtain the condition
$$m\left(\frac{d^2\varphi }{dt^2}+2g\frac{d\psi }{dt}+\frac{dg}{dt}\psi \right)\widehat{๐ฎ}๐ซ+\left(2\frac{d\varphi }{dt}+m\frac{d^2\psi }{dt^2}\right)\widehat{๐ฎ}๐ฉ=0.$$
(28)
Imposing the additional condition
$$2\frac{d\varphi }{dt}+m\frac{d^2\psi }{dt^2}=0,$$
(29)
(28) becomes
$$m\frac{d^3\psi }{dt^3}4g\frac{d\psi }{dt}2\frac{dg}{dt}\psi =0.$$
(30)
We can solve (30) thoroughly if $`g(r,t)`$ is a function of the time $`t`$ only. In this case, as before, we end up with the time-dependent isotropic harmonic oscillator. The vector $`๐ฃ`$ associated with (25) can be obtained as follows: first we integrate (29) thus obtaining
$$\varphi =\frac{m}{2}\frac{d\psi }{dt}+C,$$
(31)
where $`C`$ is an integration constant. Then making use of (25), (31) and (26) we arrive at
$`๐ฃ`$ $`=`$ $`\left[\left({\displaystyle \frac{m}{2}}{\displaystyle \frac{d\psi }{dt}}+C\right)\widehat{๐ฎ}๐ซ+\psi \widehat{๐ฎ}๐ฉ\right]๐ฉ`$
$`+\left[\left({\displaystyle \frac{m^2}{2}}{\displaystyle \frac{d^2\psi }{dt^2}}mg\psi \right)\widehat{๐ฎ}๐ซ\left({\displaystyle \frac{m}{2}}{\displaystyle \frac{d\psi }{dt}}+C\right)\widehat{๐ฎ}๐ฉ\right]๐ซ,`$
or in terms of components
$$j_i=F_{ij}u_j,$$
(33)
where $`F_{ij}`$ is defined by
$`F_{ij}`$ $`=`$ $`\left({\displaystyle \frac{m}{2}}{\displaystyle \frac{d\psi }{dt}}+C\right)p_ix_j+\psi p_ip_j`$
$`+\left({\displaystyle \frac{m^2}{2}}{\displaystyle \frac{d^2\psi }{dt^2}}mg\psi \right)x_ix_j\left({\displaystyle \frac{m}{2}}{\displaystyle \frac{d\psi }{dt}}+C\right)x_ip_j.`$
The constant $`C`$ in (4) can be made zero without loss of generality. A generalized Fradkin tensor can now be defined by
$$F_{ij}=\left(\frac{m}{2}\frac{d\psi }{dt}\right)p_ix_j+\psi p_ip_j+\left(\frac{m^2}{2}\frac{d^2\psi }{dt^2}mg\psi \right)x_ix_j\frac{m}{2}\frac{d\psi }{dt}x_ip_j.$$
(35)
From (35) we can read out the diagonal components of the generalized Fradkin tensor
$$F_{ii}=m^2\frac{d\psi }{dt}x_i\stackrel{.}{\frac{dx_i}{dt}}+\psi p_i^2+\left(\frac{m^2}{2}\frac{d^2\psi }{dt^2}mg\right)x_i^2.$$
(36)
It is not hard to see that the trace of this generalized Fradkin tensor $`๐
(๐ซ,๐ฉ,t)`$ becomes the energy of the particle when $`g(r,t)`$ is a constant field.
## 5 Obtaining explicit solutions: An alternative way
Now we wish to show how to take advantage of the vector constant $`๐ฃ`$ to obtain the solution for the Kepler and the isotropic harmonic oscillator potentials. But firstly we must establish some very general relationships between the sought for solution $`r\left(t\right)`$ and $`A(๐ซ,๐ฉ,t)`$ and $`๐ฃ`$. First notice that (10) can be recasted into the form
$$๐ฃ=mA(๐ซ,๐ฉ,t)^2\frac{d}{dt}\left[\frac{๐ซ}{A(๐ซ,๐ฉ,t)}\right],$$
(37)
where it must be kept in mind that $`A(๐ซ,๐ฉ,t)`$ satisfies (6). As we have seen before in specific examples the form of the vector constant $`๐ฃ`$ depends on the force acting on the particle. Integrating (37) we readily obtain
$$\frac{๐ซ\left(t\right)}{A(๐ซ,๐ฉ,t)}\frac{๐ซ\left(0\right)}{A(๐ซ,๐ฉ,0)}=\frac{๐ฃ}{m}_0^t\frac{d\tau }{A(๐ซ,๐ฉ,\tau )^2}.$$
(38)
Equation (38) can be given a simple but interesting geometrical interpretation. Assume that the initial conditions $`๐ซ\left(0\right)`$ and $`๐ฉ\left(0\right)`$ are known and therefore the function $`A(๐ซ,๐ฉ,0)`$ can be determined. The vector $`๐ซ\left(0\right)/A(๐ซ,๐ฉ,0)`$is therefore a constant and completely determined vector. As time increases, the R.H.S. of (38) increases. The vector on the right side of (38) though varying in time has a fixed direction which is determined by $`๐ฃ`$. Therefore, $`๐ซ\left(t\right)/A(๐ซ,๐ฉ,t)`$ must increase in order to close the triangle whose sides are the three vectors involved in (38). If the orbit is unlimited then it is easy to see that the following property ensues: there is an asymptote if in the limit $`t\mathrm{}`$ the definite integral $`_0^t\frac{dt}{A^2}`$ is constant. On the other hand, if the orbit is limited, but not necessarily closed, there will be a position vector $`๐ซ`$ whose direction is parallel to that of the vector $`๐ฃ`$ at the instant $`t^{}`$. If the length of the position vector $`๐ซ`$ is finite, we can conclude that at the same instant $`t^{}`$ the function $`A(๐ซ,๐ฉ,0)`$ must be zero. Thus, we can see that the vector $`๐ซ\left(t\right)/A(๐ซ,๐ฉ,t)`$ must be reversed at this instant and its evolution is determined by the fact that its end is on the straight line that contains $`๐ฃ`$. For $`t=t^{}+ฯต`$, where $`ฯต`$ is a positive infinitesimal number, the vector $`๐ซ\left(t^{}+ฯต\right)/A(๐ซ,๐ฉ,t^{}+ฯต)`$ changes its direction abruptly, so to speak, as shown in the figure, hence in the transition $`A(๐ซ,๐ฉ,t^{})A(๐ซ,๐ฉ,t^{}+ฯต)`$ the scalar function must change its sign.
Let us obtain the solution $`๐ซ\left(t\right)`$ for the case of the isotropic harmonic oscillator. A particular solution of (17) for $`g=\kappa ,`$ where $`\kappa `$ is the elastic cons$`\mathrm{tan}`$t is given by
$$A(t)=\mathrm{cos}(\omega t),$$
(39)
where $`\omega =\sqrt{\frac{k}{m}}`$ is the angular frequency. The solution given by (39) allows us to write
$$\frac{๐ซ\left(t\right)}{\mathrm{cos}(\omega t)}๐ซ\left(0\right)=\frac{๐ฉ\left(0\right)}{m}_0^t\frac{d\tau }{\mathrm{cos}^2(\omega \tau )}$$
(40)
The integral can be readily performed and after some simplifications we finally obtain
$$๐ซ\left(t\right)=\mathrm{cos}\omega t๐ซ\left(0\right)+\frac{1}{m\omega }\mathrm{sin}\omega t๐ฉ\left(0\right).$$
(41)
Therefore we have obtained the solution of the time-independent harmonic oscillator in an alternative way from the knowledge of the initial conditions as it should be. Notice that the general solution $`A(t)=A\left(0\right)\mathrm{cos}\left(\omega t+\theta \right)`$ would lead to the same general result. Another possible solution in the case of the time-independent isotropic harmonic oscillator is given by
$$A(๐ซ,๐ฉ,t)=\widehat{๐ฎ}๐ฉ\left(t\right),$$
(42)
as can be shown by substituting this solution into (6). This solution shows that the trajectories have no asymptotes.
Let us now show how we can obtain the orbits in the case of the time-dependent Kepler problem. Let us begin by rewriting (37) in polar coordinates on the plane. In these coordinates the angular momentum conservation law is written in the form
$$l=mr^2\frac{d\theta }{dt}$$
(43)
and this allows to rewrite (37) as
$$๐ฃ=mA^2\frac{d}{d\theta }\left(\frac{๐ซ}{A}\right)\frac{d\theta }{dt}=l\left(\frac{A}{r}\right)^2\frac{d}{d\theta }\left(\frac{๐ซ}{A}\right).$$
(44)
Introducing the unitary vectors $`\stackrel{}{๐ซ}`$ and $`\stackrel{}{\theta }`$ we can write the above equation as
$$๐ฃ=l[\frac{d}{d\theta }\left(\frac{A}{r}\right)\stackrel{}{๐ซ}+\frac{A}{r}\stackrel{}{\theta }].$$
(45)
The components of the vector $`๐ฃ`$ in the direction of $`\stackrel{}{๐ซ}`$ and $`\stackrel{}{\theta }`$ are given by
$$๐ฃ\stackrel{}{\theta }=l\frac{A}{r},$$
(46)
and
$$๐ฃ\stackrel{}{๐ซ}=l\frac{d}{d\theta }\left(\frac{A}{r}\right).$$
(47)
Equation (47) can be obtained from (46) and therefore it is redundant. In section 4 we determined a generalized Laplace-Runge-Lenz vector for the time-dependent Kepler problem. The scalar function $`A(๐ซ,๐ฉ,t)`$ associated with this vector was found to be
$$A=\varphi (t)๐ซ๐ฉm\frac{d\varphi }{dt}r^2.$$
(48)
Making use of (43) we can rewrite the linear momentum as a function of $`\theta `$ as follows
$$๐ฉ=m\frac{d๐ซ}{dt}\frac{d\theta }{dt}=\frac{l}{r^2}\frac{d๐ซ}{d\theta }.$$
(49)
Taking (49) into (48) and considering $`A`$ as a function of $`\theta `$ we obtain
$$\frac{A}{r}=l\frac{d}{d\theta }\left(\frac{\varphi }{r}\right).$$
(50)
Equations. (46) and (50) lead to
$$\frac{d}{d\theta }\left(\frac{\varphi }{r}\right)=\frac{\stackrel{}{๐ฃ\stackrel{}{\theta }}}{l^2}=\frac{j}{l^2}\mathrm{sin}(\theta \alpha ),$$
(51)
where $`\alpha `$ is the angle between $`๐ฃ`$ and the $`๐ช๐ณ`$ axis (see figure 2). In order to integrate (51) we assume that the initial conditions at $`t=0`$ are known vector functions, i.e.
$$๐ซ(0)=๐ซ_0;$$
(52)
and
$$๐ฉ(0)=๐ฉ_0.$$
(53)
In terms of polar coordinates these initial conditions are written as
$$r(\theta _0)=r_0,$$
(54)
and making use of (49)
$$\left(\frac{d๐ซ}{d\theta }\right)_{\theta =\theta _0}=\frac{r_0^2}{l}๐ฉ_0.$$
(55)
Upon integrating (51) we find
$$\frac{\varphi }{r}=\frac{\varphi _0}{r_0}+\frac{j}{l^2}\left[\mathrm{cos}(\theta _0\alpha )\mathrm{cos}(\theta \alpha )\right].$$
(56)
For the usual time-independent Kepler problem, $`\varphi =1`$ and in this case (56) takes the form
$$\frac{1}{r}=\frac{1}{r_0}+\frac{j}{l^2}\left[\mathrm{cos}(\theta _0\alpha )\mathrm{cos}(\theta \alpha )\right]$$
(57)
The scalar product between $`๐ฃ`$ as given by (16) and $`๐ซ_0`$ permit us to eliminate $`\mathrm{cos}(\theta _0\alpha )`$ and leads to the usual orbit equation
$$\frac{1}{r}=\frac{mC}{l^2}\left[1+\frac{j}{mC}\mathrm{cos}(\theta \alpha )\right].$$
(58)
If we define a new position vector $`๐ซ^{}`$ according to
$$๐ซ^{}:=\frac{๐ซ}{\varphi },$$
(59)
and redefine our time parameter according to
$$dt^{}:=\frac{dt}{\varphi ^2},$$
(60)
we can recast the equation of motion for the time-dependent Kepler problem, namely
$$m\frac{d^2๐ซ}{dt^2}=\left(\frac{m}{\varphi }\frac{d^2\varphi }{dt^2}+\frac{C}{\varphi r^3}\right)๐ซ$$
(61)
into a simpler form. According to (59) and (60) the velocity and the acceleration transform in the following way
$$\frac{d๐ซ}{dt}=๐ซ^{}\frac{d\varphi }{dt}+\frac{1}{\varphi }\frac{d๐ซ}{dt^{}}^{}$$
(62)
and
$$\frac{d^2๐ซ}{dt^2}=๐ซ^{}\frac{d^2\varphi }{dt^2}+\frac{1}{\varphi ^3}\frac{d^2๐ซ}{dt^{\mathrm{\hspace{0.17em}2}}}^{}.$$
(63)
where we have taken advantage of the fact that $`๐ซ,๐ซ^{}`$ and $`\varphi `$ can be considered as functions of $`t`$ or $`t^{}`$. Making use of (63) the equation of motion (61) can be written as
$$m\frac{d^2๐ซ}{dt^2}^{}=\frac{C}{(r^{})^3}๐ซ^{}.$$
(64)
Equation (64) corresponds to the usual time-independent Kepler problem whose solution is given by (58). Equations (59) and (60) show that the open solutions of (64) are transformed into the open solutions of (61) with the same angular size and that closed solutions of (64) are associated with spiraling solutions of (61). The period of the orbit of (64) is related to the time interval that the spiraling particle takes to cross a fixed straight line. Representing this time interval by $`T_0`$ we have
$$T=_0^{T_0}\frac{dt}{\varphi ^2}.$$
(65)
As an application of the above remarks suppose we are looking for the form of the function $`\varphi `$ which yields a circular orbit with radius $`R`$ as a solution to (61)? To find this function we see from (61) that we have to solve the following equation differential equation
$$\frac{d^2\varphi }{dt^2}+\frac{\left|C\right|}{mR^3}\varphi =\frac{\left|C\right|}{mR^3}.$$
(66)
The solution is
$$\varphi \left(t\right)=1+\varphi _0\mathrm{cos}\left(\omega t+\beta \right),$$
(67)
where $`\varphi _0`$ is a constant and $`\omega =\sqrt{\frac{\left|C\right|}{mR^3}}`$ and $`\beta `$ an arbitrary phase angle . The constant $`\varphi _0`$ may be chosen so that the transformed solution will be a given ellipse as we show below. Equation (59) leads to
$$\frac{1}{r^{}}=R^1\left[1+\varphi _0\mathrm{cos}\left(\omega t+\beta \right)\right].$$
Comparing with (58) we obtain
$$R=\frac{l^2}{m\left|C\right|},$$
(68)
and
$$\varphi _0=\frac{j}{m\left|C\right|}.$$
(69)
The period of this circular orbit is given
$$T_0=\frac{2\pi }{\omega }=2\pi \sqrt{\frac{mR^3}{\left|C\right|}}.$$
(70)
and using (65) we get
$$T=_0^{\frac{2\pi }{\omega }}\frac{dt}{(1+e\mathrm{cos}\omega t)^2}=\frac{2\pi }{\omega }\frac{1}{(1e^2)^{\frac{3}{2}}}$$
(71)
where $`e=\left|\varphi _0\right|`$ is the eccentricity. Making use of $`1e^2=R/a`$, where $`a`$ is the major semi-axis we finally obtain the orbital period
$$T=2\pi a^{\frac{3}{2}}\sqrt{\frac{m}{C}.}$$
(72)
To conclude consider the total mechanical energy associated with (64)
$$E=\frac{๐ฉ^{\mathrm{\hspace{0.17em}2}}}{2m}+\frac{C}{r^{}}=const.$$
(73)
Since $`๐ฉ^{}`$ and $`๐ฉ`$ are related by
$$๐ฉ^{}=\varphi \left(t\right)๐ฉm\dot{\varphi }\left(t\right)๐ซ$$
(74)
and $`r^{}`$ and $`r`$ by (59) we easily obtain
$$E=\varphi ^2\frac{๐ฉ^2}{2m}2\varphi \frac{d\varphi }{dt}๐ซ๐ฉ+\left(\frac{d\varphi }{dt}\right)^2\frac{r^2}{2}+C\frac{\varphi }{r}$$
(75)
which is a conserved quantity and can be interpreted as a generalization of the energy of the particle under the action of a time-dependent Kepler field.
## 6 Laplace-Runge-Lenz type of vector constants for arbitrary central fields
Equation (22) determines a time-dependent Kepler field $`g(r,t),`$ where the variables $`r`$ and $`t`$ are independent. If, however, we consider the orbit equation $`r\left(t\right)`$ we can eliminate the time variable and define the function $`g(r,t\left(r\right))`$ which can be understood as an arbitrary function of $`r.`$ For the sake of simplicity we denote this function by $`g\left(r\right)`$. The function $`\varphi \left(t\right)`$ which transforms the Kepler problem when understood as a function of $`r`$ transforms the Kepler field in an arbitrary central field. Let us write Eq. (22) in the form
$$m\frac{d^2\varphi }{dt^2}g(r,t)\varphi +\frac{C}{r^3}=0$$
(76)
and consider the transformation
$$\frac{d^2\varphi }{dt^2}=\frac{d^2\varphi }{dr^2}\left(\frac{dr}{dt}\right)^2+\frac{d\varphi }{dr}\frac{d^2r}{dt^2}$$
(77)
Energy conservation and the equation of motion allow us to write
$$\left(\frac{dr}{dt}\right)^2=\frac{2}{m}\left[EV\left(r\right)\frac{l^2}{2mr^2}\right]$$
(78)
and
$$\frac{d^2r}{dt^2}=\frac{rg\left(r\right)}{m}+\frac{l^2}{m^2r^3}$$
(79)
where $`E`$ is the energy of the particle and $`l`$ its angular momentum. Taking these three last equations into account Eq. (76) reads now
$$2\left[EV\left(r\right)\frac{l^2}{2mr^2}\right]\frac{d^2\varphi }{dr^2}+\left[rg\left(r\right)+\frac{l^2}{m^2r^3}\right]\frac{d\varphi }{dr}g\left(r\right)\varphi +\frac{C}{r^3}=0$$
(80)
Equation (80) permit us to determine the function $`\varphi \left(r\right)`$ for any potential $`V\left(r\right)`$. Thus, we conclude that a central field problem can be transformed into a time-dependent Kepler problem. When $`g\left(r\right)`$ describes the Kepler field the solution of (80) is $`\varphi \left(r\right)=1`$. Sometimes it is convenient to perform a second change of variables by defining the transformation
$$\varphi =\psi \frac{mC}{l^2}$$
(81)
Then Eq. (80) becomes
$$2\left[EV\left(r\right)\frac{l^2}{2mr^2}\right]\frac{d^2\psi }{dr^2}+\left[\frac{rg\left(r\right)}{m}+\frac{l^2}{m^2r^3}\right]\frac{d\psi }{dr}g\left(r\right)\psi =0$$
(82)
As an example consider $`V\left(r\right)=k/r`$. Then the solution of (82) is simply
$$\psi \left(r\right)=c_1\left(kmr+l^2\right)+c_2\sqrt{l^2+2kmr2mEr^2}$$
(83)
## 7 Conclusions
In this paper we have outlined a simple and effective method for treating problems related with time-dependent and time independent central force fields. In particular we have dealt with the Kepler problem and the isotropic harmonic oscillator fields. We have been able to rederive some known results from an original point of view and generalize others. The central force field has been discussed in the literature from many points of view. The difficulty in finding vector constants of motion for central fields stem from the fact that in general orbits for these type of problem are not closed, therefore any new ways to attack those problems are welcome.. In our method this difficulty is transferred, so to speak, to the obtention for each possible central field, which can be time-dependent or not, of a certain scalar function of the position, linear momentum and time. For a given central field this scalar function is a solution of (6). In the general case, the obtention of the scalar function is a difficult task. Judicious guesses, however, facilitate the search for solution of (6) and this is what we have done here. |
warning/0003/hep-th0003286.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Much progress has been achieved recently in understanding the full non-linear structure of certain Kaluza-Klein sphere reductions. To be specific, we have in mind the remarkable cases where it is consistent to include lower-dimensional fields in the reduction Ansatz that parameterise inhomogeneous deformations of the internal sphere metric. Generically, one would expect that performing a truncation of the complete Kaluza-Klein towers of massless and massive modes to the purely massless sector would give rise to inconsistencies beyond the linearised level, since currents built from the massless fields would act as sources for the massive fields that have been set to zero. Indeed this is exactly what usually happens; one cannot make a consistent Kaluza-Klein $`n`$-sphere reduction of a generic theory in which all the massless fields, including, in particular, the full set of $`SO(n+1)`$ gauge fields are retained. However, in certain very special cases a consistent reduction of this kind is possible.
An important early example of this type was much studied in the 1980โs, namely the seven-sphere compactification of eleven-dimensional supergravity. It was first shown at the level of linearised fluctuations around the AdS$`{}_{4}{}^{}\times S^7`$ Freund-Rubin vacuum solution that the massless modes described four-dimensional $`N=8`$ gauged $`SO(8)`$ supergravity . Subsequently, it was shown that despite all the apparent obstacles, the reduction to the massless $`N=8`$ multiplet can be carried through as an exact embedding at the full non-linear level , although the construction is an extremely complex one. It has long been believed that consistent reductions should also be possible in the case of the $`S^5`$ compactification of type IIB supergravity, and the $`S^4`$ compactification of eleven-dimensional supergravity, to yield the maximal gauged $`SO(6)`$ supergravity in $`D=5`$, and the maximal gauged $`SO(5)`$ supergravity in $`D=7`$, respectively. Indeed, the consistent $`S^4`$ reduction Ansatz from $`D=11`$ has recently been constructed . No analogous construction exists for the complete massless reduction on $`S^5`$.
It is sometimes helpful to study sphere reductions in which only a subset of the complete set of massless fields is retained, in such a way that one still has the non-triviality of the inhomogeneous sphere deformations, while at the same time making the problem of obtaining an explicit reduction Ansatz more tractable and manageable. This can be especially useful if one actually wants to use the Ansatz for the purpose of lifting solutions of the lower-dimensional theory back to the higher dimension, in which case full explicit reduction formulae are highly advantageous. In this spirit, consistent reductions in the three cases mentioned above were constructed for truncations in which only the maximal abelian subgroups $`U(1)^4`$, $`U(1)^3`$ and $`U(1)^2`$ of the full $`SO(8)`$, $`SO(6)`$ and $`SO(5)`$ gauge groups were retained, together with associated scalar fields . The $`U(1)^3`$ example provided the first concrete evidence for the consistency of the $`S^5`$ reduction of type IIB supergravity. The reduction Ansรคtze were sufficiently explicit that they could be used for the purpose of lifting certain AdS black-hole solutions back to the higher dimension , where they become rotating D3-branes and M-branes .
Other consistent reductions involving subsets of the complete massless sector have subsequently been constructed, including an $`S^4`$ reduction to give $`N=2`$ gauged $`SU(2)`$ supergravity in $`D=7`$ ; an $`S^5`$ reduction to $`N=4`$ gauged $`SU(2)\times U(1)`$ supergravity in $`D=5`$ ; and an $`S^7`$ reduction to $`N=4`$ gauged $`SO(4)`$ supergravity in $`D=4`$ . In addition, the $`N=2`$ gauged $`SU(2)`$ supergravity in $`D=6`$ was obtained via a consistent reduction from massive type IIA supergravity on a locally $`S^4`$ internal space . This is actually the largest possible supersymmetry for a gauged theory in $`D=6`$, even though the maximum supersymmetry for ungauged supergravity is $`N=4`$.
Somtimes, It can also be useful to construct a Kaluza-Klein sphere reduction in which a non-supersymmetric truncation of the massless supermultiplet is made. One example of this type involved truncating the maximal supergravities in $`D=4`$, 5 and 7 to a subsector comprising just gravity and a set of 7, 5 or 4 scalars respectively. These scalars correspond to the diagonal subset of fields in the unimodular symmetric tensors $`T_{ij}`$ describing the scalars in the $`SL(8,\text{I}\mathrm{R})/SO(8)`$, $`SL(6,\text{I}\mathrm{R})/SO(6)`$ and $`SL(5,\text{I}\mathrm{R})/SO(5)`$ scalar submanifolds of the full supergravities. In , the full non-linear reduction Ansรคtze for these embedding were constructed, and proved to be consistent.
Another example of a non-supersymmetric truncation was constructed in , where the full set of twenty scalars $`T_{ij}`$ of the coset $`SL(6,\text{I}\mathrm{R})/SO(6)`$ were retained in an $`S^5`$ reduction from $`D=10`$. Consistency now requires that one include also the full set of $`SO(6)`$ Yang-Mills gauge fields. In fact only the metric and self-dual 5-form of the type IIB supergravity are involved in this reduction, and it can equivalently be viewed as a Kaluza-Klein reduction of a theory of pure gravity plus self-dual 5-form in $`D=10`$, with all massless fields retained in $`D=5`$. (The truncation of type IIB supergravity to just gravity and the self-dual 5-form is itself a consistent one in $`D=10`$.) The self-duality of the 5-form is crucial for the consistency of the reduction.
One should not conclude from the listing of examples above that consistent Kaluza-Klein sphere reductions are a commonplace. In fact, if we restrict attention to cases where one starts in the higher dimension with just gravity and a $`p`$-form field strength, then it turns out that the only cases that can give consistent reductions are related to the examples mentioned above.<sup>1</sup><sup>1</sup>1To be precise, we should emphasise that what we are discussing here is cases where all of the $`SO(n+1)`$ Yang-Mills gauge fields associated with the isometry group of the $`n`$-sphere are retained in the truncation, together with other associated massless scalars. The reason for this can be understood as follows. For reductions of the type we are considering, where the lower-dimensional theory obtained by the $`S^n`$ reduction has an $`SO(n+1)`$ local gauge symmetry, it is essential that the ungauged theory that would result from performing a reduction on the $`n`$-torus rather than the $`n`$-sphere should have scalars described by a coset manifold $`G/H`$ such that $`H`$ at least contains $`SO(n+1)`$. The reason for this is that in the process of gauging the ungauged theory, a subgroup $`SO(n+1)`$ of the global symmetry $`G`$ must become local, and this subgroup must be contained within $`H`$. Now if a generic theory of gravity and antisymmetric tensors is reduced on $`T^n`$, it will give rise to a lower-dimensional theory with a $`GL(n,\text{I}\mathrm{R})`$ global symmetry , for which the maximal compact subgroup is $`SO(n)`$. This is insufficient for allowing an $`SO(n+1)`$ gauging. Note that in particular this argument shows that it is not possible to perform a consistent $`n`$-sphere reduction of a pure gravity theory, in which the Yang-Mills fields of $`SO(n+1)`$ are retained.
In certain very special theories, the $`GL(n,\text{I}\mathrm{R})`$ global symmetry arising from a $`T^n`$ reduction is enhanced to a larger symmetry, as a result of โconspiraciesโ between scalars coming from the Kaluza-Klein reduction of the metric and of the other higher-dimensional fields. However, as discussed in , such cases are very few and far between. In particular, if we consider a $`D`$-dimensional theory consisting of gravity and a single $`p`$-form field strength, an enhancement of the global symmetry can occur only if $`(D,p)`$ is equal to $`(11,4)`$, $`(11,7)`$ or $`(10,5)`$. Since a 7-form in $`D=11`$ is dual to a 4-form this means that the only cases with symmetry enhancements are associated with $`D=11`$ supergravity and type IIB supergravity. The corresponding enhanced symmetries in each case are to $`SL(8,\text{I}\mathrm{R})/SO(8)`$ in $`D=4`$, $`SL(5,\text{I}\mathrm{R})/SO(5)`$ in $`D=7`$, and $`SL(6,\text{I}\mathrm{R})/SO(6)`$ in $`D=5`$.<sup>2</sup><sup>2</sup>2Actually further conditions must be fulfilled in order for the symmetry enhancements to take place. In the case of $`D=11`$ reduced to $`D=4`$ on $`T^7`$, the enhancement of $`GL(7,\text{I}\mathrm{R})`$ to $`SL(8,\text{I}\mathrm{R})`$ requires that the 4-form have an $`FFA`$ term in $`D=11`$ with precisely the coefficient dictated by supersymmetry; this in fact means that the enhanced symmetry is even larger, namely $`E_7`$. For the $`D=10`$ theory reduced on $`T^5`$, the enhancement from $`GL(5,\text{I}\mathrm{R})`$ to $`SL(6,\text{I}\mathrm{R})`$ requires that the 5-form be self-dual (or anti-self-dual). These enhancements then allow the $`SO(8)`$, $`SO(6)`$ and $`SO(5)`$ gaugings, respectively.<sup>3</sup><sup>3</sup>3The importance of enhancements of the global symmetry in toroidal reductions was also observed in , although it was assumed there that the phenomenon was much more widespread than is actually the case.
Note that whereas there is a consistent $`S^5`$ reduction of gravity with a self-dual 5-form in which only the metric, the $`SO(6)`$ Yang-Mills fields $`A_{\left(1\right)}^{ij}`$ and the 20 scalars $`T_{ij}`$ are retained, the situation is a little different in the $`S^4`$ and $`S^7`$ reductions from $`D=11`$. In addition to keeping the corresponding Yang-Mills fields $`A_{\left(1\right)}^{ij}`$ and scalars $`T_{ij}`$, the consistency of the $`S^4`$ reduction requires also keeping the five 3-forms $`A_{\left(3\right)}^i`$ of the seven-dimensional theory, whilst the $`S^7`$ reduction instead requires also keeping the 35 pseudoscalars $`\varphi _{[ijk\mathrm{}]_+}`$ (self-dual in the $`SO(8)`$ indices). These additional fields are needed in the reductions because the Yang-Mills fields act as sources for them . In fact we can summarise the situation in all three of these examples of the $`S^4`$, $`S^7`$ and $`S^5`$ reductions as follows. In all cases, the consistent $`n`$-sphere reduction that includes all the Yang-Mills fields of $`SO(n+1)`$ requires one to include all the massless fields in the lower-dimensional theory. Thus in the $`S^5`$ case, if we reduce the theory of gravity and the self-dual 5-form then the metric, the $`SO(6)`$ Yang-Mills fields and the scalars $`T_{ij}`$ indeed constitute the complete set of massless fields in five dimensions. In the $`S^4`$ reduction from $`D=11`$ the five 3-forms $`A_{\left(3\right)}^i`$ are massless too, and indeed they must be included also in the consistent reduction. Likewise, in the $`S^7`$ reduction from $`D=11`$ the 35 pseudoscalars $`\varphi _{[ijk\mathrm{}]_+}`$ are also massless, and indeed they must be included in the consistent reduction.
Further possibilities for consistent $`n`$-sphere reductions in which all the $`SO(n+1)`$ Yang-Mills fields are retained can arise if we consider a slightly enlarged higher-dimensional theory, now with a dilatonic scalar as well as gravity and the $`p`$-form field strength. Again, the key point is that an enhancement of the $`GL(n,\text{I}\mathrm{R})`$ global symmetry that would occur for the reduction of a generic theory on $`T^n`$ is needed, in order that the scalar coset manifold in the lower dimension should have a denominator group that is large enough to contain the desired $`SO(n+1)`$ local symmetry group of the theory reduced on $`S^n`$. It turns out that by including the dilatonic scalar in the higher dimension, the necessary symmetry enhancements can be achieved when the $`p`$-form field strength is either a 2-form, or a 3-form. This opens up the possibility of finding consistent Kaluza-Klein reductions on $`S^2`$, for the 2-form case, and on $`S^3`$ or on $`S^{D3}`$, for the 3-form case. These, together with the previous $`D=11`$ and $`D=10`$ examples, would then constitute the complete list of possibilities for consistent Kaluza-Klein $`n`$-sphere reductions within the class of theories we are considering, in which all the Yang-Mills fields of $`SO(n+1)`$ are retained.
In this paper, we construct the complete and explicit non-linear Kaluza-Klein Ansรคtze for these three new possibilities. We begin in section 2 with a detailed discussion of the global symmetry enhancements that can occur in the toroidal reductions of theories with gravity, a $`p`$-form field strength and a dilaton, in order to establish what are the possibilities for consistent sphere reduction. In section 3 we construct the Ansatz for the consistent reduction of gravity plus a 3-form and a dilaton on $`S^3`$, keeping all the gauge fields of $`SO(4)`$ and the ten scalars of the symmetric tensor $`T_{ij}`$, together with the 2-form potential $`A_{\left(2\right)}`$. In section 4, we examine two truncations of the $`S^3`$ reduction, in which only certain subsets of the massless fields are retained, in order to make contact with previous results in the literature. In section 5 we consider the โdualโ of the $`S^3`$ reduction namely, the reduction instead on $`S^{D3}`$. Again, we find a consistent reduction Ansatz, in which all the gauge fields of $`SO(D2)`$ are retained, together with the $`\frac{1}{2}(D1)(D2)`$ scalars in $`T_{ij}`$. A case of particular interest is the $`S^7`$ reduction from $`D=10`$, since then the resulting three-dimensional theory is the bosonic sector of a gauged $`SO(8)`$ supergravity, of a type not previously constructed in the literature. In section 6 we construct the consistent Kaluza-Klein Ansatz for the reduction of a theory of gravity, a dilaton and a 2-form field strength on $`S^2`$. In this case the Ansatz includes all three gauge fields of $`SO(3)`$, together with the six scalars in $`T_{ij}`$.
Note that in two of the new cases that we consider here, namely the $`S^2`$ reduction of the theory with a 2-form field strength, and the $`S^{D3}`$ reduction of the theory with a 3-form field strength, the totality of massless fields in the Kaluza-Klein reduced lower-dimensional theories comprise the metric, the Yang-Mills gauge fields $`A_{\left(1\right)}^{ij}`$, and the scalars $`T_{ij}`$. Thus these new examples of consistent sphere reductions are akin to the $`S^5`$ reduction of gravity plus a self-dual 5-form, in that no additional massless fields are present that must also be included in the reduction Ansatz. By constrast, in the new $`S^3`$ reduction that we construct here, we must additionally include the 2-form potential $`A_{\left(2\right)}`$ in the Ansatz. This is similar to the situation for the $`S^4`$ and $`S^7`$ reductions, where, as we discussed previously, additional massless fields are present, and must be included, for consistency.
## 2 Possibilities for $`SO(n+1)`$ Kaluza-Klein reductions on $`S^n`$
As we mentioned in the Introduction, a consistent Kaluza-Klein reduction on $`S^n`$ that retains all the gauge fields of $`SO(n+1)`$ will be possible only if there is a suitable enhancement of the generic $`GL(n,\text{I}\mathrm{R})`$ global symmetry that arises instead in a $`T^n`$ reduction, so that the denominator group of the generic scalar manifold $`GL(n,\text{I}\mathrm{R})/O(n)`$ becomes large enough to contain $`SO(n+1)`$. We also mentioned that there are only rather limited circumstances under which these symmetry enhancements can occur.
The reason why the possibilities for symmetry enhancements are so restrictive is discussed extensively in . The scalars divide into $`n`$ โdilatons,โ $`\stackrel{}{\varphi }`$ coming from the diagonal metric components of the internal $`n`$-torus, with the rest being โaxionsโ $`\chi _i`$ coming from the off-diagonal metric components and the reduction of the antisymmetric tensor. Each dilaton has a kinetic term of the form $`\frac{1}{2}e^{\stackrel{}{c}_i\stackrel{}{\varphi }}(\chi _i)^2`$, where $`\stackrel{}{c}_i`$ is the associated constant โdilaton vectorโ characterising the coupling of the dilatons to that particular axion.
In the $`n`$-torus reduction of a theory of $`D`$-dimensional gravity plus $`p`$-form field strength with general values of $`D`$ and $`p`$, the global symmetry will be $`GL(n,\text{I}\mathrm{R})`$. In fact the dilaton vectors $`\stackrel{}{b}_i`$ associated with the axions coming from the metric form the complete set of positive roots of the $`SL(n,\text{I}\mathrm{R})`$ algebra . The dilaton vectors $`\stackrel{}{a}_i`$ associated with the axions coming from the $`p`$-form field strength then form the weights of some representation under $`SL(n,\text{I}\mathrm{R})`$. If an enhancement of the global symmetry is to occur, it must be that some or all of the dilaton vectors $`\stackrel{}{a}_i`$ โconspireโ to become the additional positive roots of the enhanced symmetry algebra. However, this cannot occur in general, because the lengths of vectors $`\stackrel{}{a}_i`$ coming from the $`p`$-form will be incommensurate with the lengths of the vectors $`\stackrel{}{b}_i`$ coming from the metric.<sup>4</sup><sup>4</sup>4We first, of course, establish a canonical normalisation for the dilaton kinetic terms, so that comparisons of the lengths are meaningful.
A convenient way to characterise the lengths of the various dilaton vectors was introduced in . Rather than using the quantity $`|\stackrel{}{c}|^2`$ itself, it is convenient to introduce a constant $`\delta `$, related to $`|\stackrel{}{c}|^2`$ by
$$|\stackrel{}{c}|^2=\delta \frac{2(m1)(Dm1)}{D2},$$
(1)
where $`D`$ is the spacetime dimension, and $`m`$ is the degree of the field-strength whose dilaton coupling is $`e^{\stackrel{}{c}\stackrel{}{\varphi }}`$. (Note that the all the field strengths in the $`D=11`$ and $`D=10`$ supergravities have $`\delta =4`$ couplings .) The key point about this parameterisation is that $`\delta `$ is preserved under toroidal Kaluza-Klein reduction. This makes it rather easy to see when the possibility of an enhancement of the global symmetry can occur. First, we note that the dilaton vectors $`\stackrel{}{b}_i`$ associated with the axions coming from the Kaluza-Klein reduction of the metric always have $`\delta =4`$. It follows therefore that if the dilaton vectors $`\stackrel{}{a}_i`$ associated with axions coming from the $`p`$-form field strength in the higher dimension are to have the same lengths as the $`\stackrel{}{b}_i`$, then the coupling of the field strength must also have $`\delta =4`$.<sup>5</sup><sup>5</sup>5Since all the $`\stackrel{}{a}_i`$ themselves have equal length, and all the $`\stackrel{}{b}_i`$ themselves have equal length, it follows that to get a simply-laced enhanced symmetry group we would need that the length of the $`\stackrel{}{a}_i`$ and the length of the $`\stackrel{}{b}_i`$ should be equal. This turns out to be the only situation where relevant symmetry enhancements occur, within the framework of the higher-dimensional Lagrangians (3). Thus although the case $`D=6`$ with a $`\delta =2`$ self-dual 3-form gives an enhancement to $`O(3,4)`$ after a $`T^3`$ reduction to $`D=3`$ , and $`D=5`$ with a $`\delta =\frac{4}{3}`$ 2-form gives an enhancement to $`G_2`$ after a $`T^2`$ reduction to $`D=3`$ , neither of these non-simply-laced cases would seem to indicate the possibility of consistent $`S^3`$ or $`S^2`$ reductions. If we are starting in $`D`$ dimensions with a theory with just gravity and the $`p`$-form field strength, but no dilaton, this means that in $`D`$ dimensions we have $`\stackrel{}{c}=0`$, and $`m=p`$, and so to have $`\delta =4`$ we must have
$$2D4=(p1)(Dp1).$$
(2)
It is easily verified that the only integer solutions are $`(D,p)=(11,4)`$, $`(11,7)`$ and $`(10,5)`$.
The possibilities for achieving the necessary enhancement of the global $`GL(n,\text{I}\mathrm{R})`$ symmetry can be broadened considerably if a dilatonic scalar is included in the original theory in $`D`$ dimensions, since now there is the possibility of adjusting its coupling to the $`p`$-form field strength so that the corresponding value of $`\delta `$ is equal to 4. Thus we may consider the $`D`$-dimensional Lagrangian
$$_D=\widehat{R}\widehat{}\text{1}\mathrm{l}\frac{1}{2}\widehat{}d\widehat{\varphi }d\widehat{\varphi }\frac{1}{2}e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(p\right)}\widehat{F}_{\left(p\right)},$$
(3)
with $`a`$ chosen so that
$$a^2=4\frac{2(p1)(Dp1)}{D2}.$$
(4)
Note that we put hats on all the fields in (3), to indicate that they are higher-dimensional quantities.
The first point to notice is that the requirement that the constant $`a`$ should be real<sup>6</sup><sup>6</sup>6One might in principle consider also the possibility that $`a`$ could be imaginary. This would be equivalent to having a ghost-like kinetic term for the dilaton in the $`D`$-dimensional theory. This could not lead to any useful global symmetry enhancements from the point of view of sphere reductions that retain the $`SO(n+1)`$ gauge fields. There might be possible implications for consistent reductions on spaces with non-compact symmetry groups. is a rather restrictive one, since it implies
$$p^23Dp+3D50.$$
(5)
Taken together with the fact that obviously $`p`$ cannot exceed $`D`$, this implies that the only additional possibilities opened up by the inclusion of the dilaton are for $`p=2`$, 3, $`(D2)`$ or $`(D3)`$. The last two here are just the Hodge duals of $`p=2`$ and $`p=3`$, so we need not consider them as distinct cases. For $`p=2`$ and $`p=3`$, the relation (4) gives
$`p=2:`$ $`a^2={\displaystyle \frac{2(D1)}{D2}},`$ (6)
$`p=3:`$ $`a^2={\displaystyle \frac{8}{D2}}.`$ (7)
Thus we see that if we start in $`D`$ dimensions with the Lagrangian (3) with a 2-form or 3-form field strength, we can achieve a $`\delta =4`$ dilaton coupling in any dimension $`D`$, and thus we can expect to find an enhancement of the $`GL(n,\text{I}\mathrm{R})`$ global symmetry after dimensional reduction on $`T^n`$. Indeed this is the case.
First, let us consider the case $`p=3`$, where we make a $`T^n`$ reduction of (3) with $`a`$ given by (7). The global symmetry is indeed enhanced, and the scalar manifold in $`(Dn)`$ dimensions will be
$`Dn>3:`$ $`\text{I}\mathrm{R}\times {\displaystyle \frac{O(n,n)}{O(n)\times O(n)}},`$ (8)
$`Dn=3:`$ $`{\displaystyle \frac{O(D2,D2)}{O(D2)\times O(D2)}}.`$ (10)
This $`p=3`$ case corresponds precisely to the T-duality symmetry of the toroidally-reduced bosonic string. Note that if $`Dn=3`$ the usual T-duality group $`O(D3,D3)`$ of the string theory reduced on $`T^n`$ is further enhanced to the non-perturbative U-duality group $`O(D2,D2)`$. Using the 3-form field strength, we can then consider either an $`S^3`$ or an $`S^{D3}`$ Kaluza-Klein reduction.
If we take $`n=3`$, we see that the the scalar coset manifold from a $`T^3`$ reduction will be
$$\text{I}\mathrm{R}\times \frac{O(3,3)}{O(3)\times O(3)}\text{I}\mathrm{R}\times \frac{SL(4,\text{I}\mathrm{R})}{SO(4)}.$$
(11)
There will also be six gauge potentials coming from the Kaluza-Klein reduction on $`T^3`$. This implies that the $`SO(4)`$ subgroup of the $`SL(4,\text{I}\mathrm{R})`$ global symmetry group can be gauged, with the six vector potentials becoming the Yang-Mills fields of $`SO(4)`$. It is then natural to conjecture that this gauged theory may be directly obtainable as a Kaluza-Klein reduction on $`S^3`$. It is far from obvious that such a reduction would be consistent, since unlike the toroidal reduction there is no obvious group-theoretic argument that would guarantee the consistency at the non-linear level.<sup>7</sup><sup>7</sup>7If we were gauging only the left-acting $`SU(2)`$ or only the right-acting $`SU(2)`$ of the $`SO(4)SU(2)_L\times SU(2)_R`$ isometry of the 3-sphere (which is itself the group manifold $`SU(2)`$), then the consistency would be guaranteed, since the retained fields would then all be singlets under the other $`SU(2)`$, but this is no longer the case when the gauge fields of the full isometry group are retained. In fact we shall discuss the truncation to a single $`SU(2)`$ subgroup in section 4. In the next section, we shall explicitly show that the reduction on $`S^3`$, in which the full set of $`SO(4)`$ gauge fields are retained, is in fact consistent at the full non-linear level.
Now let us consider instead the $`T^{D3}`$ reduction of (3), again with $`p=3`$. The reduced theory will now be in three dimensions, and the scalar coset manifold will be given by (10), provided that $`a`$ satisfies (7). Note that the further symmetry enhancement of this $`Dn=3`$ case occurs because the complete field content of the resulting three-dimensional theory (except for the metric) can be described by scalars, since in three dimensions one can dualise all the vector potentials to scalars. The coset (10) can also be described as
$$\frac{GL(D2,\text{I}\mathrm{R})}{O(D2)}V,$$
(12)
where $`V`$ is an irreducible representation under $`GL(D2,\text{I}\mathrm{R})`$ of dimension $`\frac{1}{2}(D2)(D3)`$; this is the same as the dimension of the adjoint representation of $`O(D2)`$.
The scalars in the representation $`V`$ can be dualised to vector potentials,<sup>8</sup><sup>8</sup>8The description (12) would arise naturally if one dualised the $`(D3)`$ vector potentials coming from the direct reduction of the original potential $`\widehat{A}_{\left(2\right)}`$, but left all other vector potentials (including the Kaluza-Klein vectors) in their original undualised forms. suggesting that the $`O(D2)`$ denominator group in (12) can be gauged. Then we may conjecture that this gauged three-dimensional theory can alternatively be obtained as a reduction of the original $`D`$-dimensional theory on the sphere $`S^{D3}`$. In section 5, we shall demonstrate that there is indeed such a consistent reduction on $`S^{D3}`$, in which all the gauge fields $`A_{\left(1\right)}^{ij}`$ of $`O(D2)`$ are retained, together with scalars described by the symmetric tensor $`T_{ij}`$, where $`i`$ is a vector index of $`O(D2)`$.
Finally, let us consider the Lagrangian (3) with $`p=2`$, where the dilaton coupling for the 2-form is given by (6). The Lagrangian (3) is then in fact precisely the $`S^1`$ dimensional reduction of pure gravity in $`D+1`$ dimensions. Consequently, the scalar manifold in $`(Dn)`$ dimensions after reducing (3) on $`T^n`$ will be enhanced to
$$\frac{GL(n+1,\text{I}\mathrm{R})}{SO(n+1)}.$$
(13)
With a 2-form field strength we have in principle two possibilities for sphere reductions, namely on $`S^2`$ or on $`S^{D2}`$. The latter would be somewhat degenerate, since the lower-dimensional theory would be in $`D=2`$, so we shall just consider the $`S^2`$ possibility here. If we take $`n=2`$, the denominator group in (13) is exactly what is needed to allow an $`SO(3)`$ gauging. We may then conjecture that this gauged theory should alternatively be directly obtainable as a consistent Kaluza-Klein reduction on $`S^2`$, keeping all three of the $`SO(3)`$ Yang-Mills gauge fields, together with six scalars described by the symmetric tensor $`T_{ij}`$. We shall in fact construct this consistent reduction in section 6.
We conclude this section with a Table that summarises all the cases where consistent sphere reductions of a $`D`$-dimensional theory comprising gravity, a $`p`$-form field strength and (in some cases) a dilatonic scalar, are possible. In all cases, we are concerned with the situation where all the Yang-Mills fields of the $`SO(n+1)`$ isometry group of the $`n`$-sphere can be included in the reduction Ansatz.
$`p`$-form Dilaton Higher-Dim Lower-Dim. Sphere Gauge Group Extra fields $`F_{\left(2\right)}`$ Yes Any $`D`$ $`D2`$ $`S^2`$ $`SO(3)`$ None $`F_{\left(3\right)}`$ Yes Any $`D`$ $`D3`$ $`S^3`$ $`SO(4)`$ $`A_{\left(2\right)}`$ $`F_{\left(3\right)}`$ Yes Any $`D`$ $`3`$ $`S^{D3}`$ $`SO(D2)`$ None $`F_{\left(4\right)}`$ No 11 7 $`S^4`$ $`SO(5)`$ $`A_{\left(3\right)}^i`$ $`F_{\left(4\right)}`$ No 11 4 $`S^7`$ $`SO(8)`$ $`\varphi _{[ijk\mathrm{}]_+}`$ $`F_{\left(5\right)}=F_{\left(5\right)}`$ No 10 5 $`S^5`$ $`SO(6)`$ None
Table 1: The possible cases for Kaluza-Klein $`S^n`$ reduction with $`SO(n+1)`$ gauge fields. The last column indicates what additional fields, beyond the metric, the gauge fields $`A_{\left(1\right)}^{ij}`$ and the scalars $`T_{ij}`$, are massless, and must therefore be included, in a consistent truncation (see discussion in section 1).
## 3 Consistent $`S^3`$ reduction
We start from the bosonic string in $`D`$ dimensions, with the low-energy effective Lagrangian<sup>9</sup><sup>9</sup>9Later, in section 7, we shall include the cosmological type term $`\frac{1}{2}m^2(D26)e^{\frac{1}{2}a\widehat{\varphi }}`$ that arises when $`D26`$, as a result of the conformal anomaly. For now, we restrict attention to the purely classical Lagrangian for gravity coupled to a 3-form and a dilaton.
$$_D=\widehat{R}\widehat{}\text{1}\mathrm{l}\frac{1}{2}\widehat{}d\widehat{\varphi }d\widehat{\varphi }\frac{1}{2}e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(3\right)}\widehat{F}_{\left(3\right)},$$
(14)
where the positive constant $`a`$ is given by (7) so that the global symmetry from a $`T^n`$ reduction would be $`\text{I}\mathrm{R}\times O(n,n)`$ rather than merely $`GL(n,\text{I}\mathrm{R})`$, as discussed in section 2. As we argued there, we can now conjecture that it should be possible to perform a consistent Kaluza-Klein reduction on $`S^3`$, keeping all the $`SO(4)`$ Yang-Mills fields $`A_{\left(1\right)}^{ij}`$, together with the scalar fields described by the symmetric tensor $`T_{ij}`$, where $`i`$ is a vector under $`SO(4)`$, and also the 2-form potential $`A_{\left(2\right)}`$.
We find that there is indeed an Ansatz for a consistent Kaluza-Klein reduction on $`S^3`$, given by
$`d\widehat{s}_D^2=Y^{\frac{1}{D2}}\left(\mathrm{\Delta }^{\frac{2}{D2}}ds_{D3}^2+g^2\mathrm{\Delta }^{\frac{D4}{D2}}T_{ij}^1๐\mu ^i๐\mu ^j\right),`$
$`e^{\sqrt{(D2)/2}\widehat{\varphi }}=\mathrm{\Delta }^1Y^{(D4)/4},`$
$`\widehat{F}_{\left(3\right)}=F_{\left(3\right)}+\frac{1}{6}ฯต_{i_1i_2i_3i_4}(g^2U\mathrm{\Delta }^2๐\mu ^{i_1}๐\mu ^{i_2}๐\mu ^{i_3}\mu ^{i_4}`$ (15)
$`3g^2\mathrm{\Delta }^2D\mu ^{i_1}๐\mu ^{i_2}๐T_{i_3j}T_{i_4k}\mu ^j\mu ^k3g^1\mathrm{\Delta }^1F_{\left(2\right)}^{i_1i_2}๐\mu ^{i_3}T_{i_4j}\mu ^j),`$
where
$`\mu ^i\mu ^i=1,\mathrm{\Delta }=T_{ij}\mu ^i\mu ^j,U=2T_{ik}T_{jk}\mu ^i\mu ^j\mathrm{\Delta }T_{ii},`$
$`Y=det(T_{ij}),`$ (16)
and the indices $`i,j,\mathrm{}`$ range of 4 values. Here, and in the rest of the paper, a summation over repeated $`SO(n+1)`$ indices is understood. The gauge-covariant exterior derivative $`D`$ is defined so that
$$๐\mu ^i=d\mu ^i+gA_{\left(1\right)}^{ij}\mu ^j,๐T_{ij}=dT_{ij}+gA_{\left(1\right)}^{ik}T_{kj}+gA_{\left(1\right)}^{jk}T_{ik},$$
(17)
where $`A_{\left(1\right)}^{ij}`$ denotes the $`SO(4)`$ gauge potentials coming from the isometry group of the 3-sphere, and
$$F_{\left(2\right)}^{ij}=dA_{\left(1\right)}^{ij}+gA_{\left(1\right)}^{ik}A_{\left(1\right)}^{kj}.$$
(18)
Thus the lower-dimensional fields appearing in the Kaluza-Klein Ansatz comprise the metric $`ds_{D3}^2`$, the six gauge potentials $`A_{\left(1\right)}^{ij}`$ of $`SO(4)`$, the ten scalar fields described by the symmetric tensor $`T_{ij}`$, and the 2-form potential $`A_{\left(2\right)}`$, whose (Chern-Simons modified) field strength is $`F_{\left(3\right)}`$. The form of the Ansatz that we have obtained here closely parallels the structure found in for the $`S^5`$ reduction of type IIB supergravity.
In order to demonstrate the consistency of the Kaluza-Klein reduction with the above Ansatz, we substitute it into the $`D`$-dimensional equations of motion<sup>10</sup><sup>10</sup>10We shall not consider the reduction of the $`D`$-dimensional Einstein equation in detail in this paper, on account of its complexity; this will be addressed in future work. In practice, in all cases that have been examined, the Einstein equation seems always to give consistent results provided that the equations of motion for all the other fields are consistent. Furthermore, the agreement of our reduction Anstaz with previously-established special cases provides further supporting evidence for the consistency of the Einstein equation. that follow from (14). These are
$`d\widehat{}d\widehat{\varphi }`$ $`=`$ $`\frac{1}{2}ae^{a\varphi }\widehat{F}_{\left(3\right)}\widehat{}\widehat{F}_{\left(3\right)},`$
$`d(e^{a\varphi }\widehat{}\widehat{F}_{\left(3\right)})`$ $`=`$ $`0,`$ (19)
$`\widehat{R}_{MN}`$ $`=`$ $`\frac{1}{2}_M\varphi _N\varphi +\frac{1}{4}\left(\widehat{F}_{MN}^2{\displaystyle \frac{2}{3(D2)}}\widehat{F}_{\left(3\right)}^2\widehat{g}_{MN}\right).`$
In addition, we have the Bianchi identity $`d\widehat{F}_{\left(3\right)}=0`$. Taking this first, we find upon substituting $`\widehat{F}_{\left(3\right)}`$ from (15) into $`d\widehat{F}_{\left(3\right)}=0`$ that the lower-dimensional field $`F_{\left(3\right)}`$ must satisfy the Bianchi identity
$$dF_{\left(3\right)}=\frac{1}{8}ฯต_{i_1i_2i_3i_4}F_{\left(2\right)}^{i_1i_2}F_{\left(2\right)}^{i_3i_4}.$$
(20)
All other terms arising from $`d\widehat{F}_{\left(3\right)}`$ vanish identically. (The calculation is quite involved, and the Schoutens identity $`ฯต_{[i_1i_2i_3i_4}V_{i_5]}0`$ is useful.)
In order to substitute the Ansatz into the field equation for $`\widehat{F}_{\left(3\right)}`$, we must first calculate the Hodge dual of $`\widehat{F}_{\left(3\right)}`$ given in (15). This is a straightforward, although somewhat involved calculation, and we find
$`e^{\sqrt{8/(D2)}\widehat{\varphi }}\widehat{}\widehat{F}_{\left(3\right)}`$ $`=`$ $`{\displaystyle \frac{1}{6g^3}}ฯต_{ijk\mathrm{}}Y^1F_{\left(3\right)}\mu ^i๐\mu ^j๐\mu ^k๐\mu ^{\mathrm{}}gUฯต_{D3}`$
$`+g^1T_{ij}^1๐T_{jk}(\mu ^k๐\mu ^i){\displaystyle \frac{1}{2g^2}}T_{ik}^1T_{jl}^1F_{\left(2\right)}^{ij}๐\mu ^k๐\mu ^{\mathrm{}}.`$
Substituting into (19), with $`a`$ given by (7), we (eventually) read off the lower-dimensional equations of motion
$`(1)^D๐(T_{ik}^1T_j\mathrm{}^1F_{\left(2\right)}^k\mathrm{})=2gT_{k[i}^1๐T_{j]k}\frac{1}{2}ฯต_{ijk\mathrm{}}Y^1F_{\left(3\right)}F_{\left(2\right)}^k\mathrm{},`$
$`(1)^D๐(\stackrel{~}{T}_{ik}^1๐\stackrel{~}{T}_{kj})=2g^2[2T_{ik}T_{jk}T_{ij}T_{kk}]ฯต_{D3}T_\mathrm{}m^1T_{ik}^1F_{\left(2\right)}^\mathrm{}kF_{\left(2\right)}^{mj}`$
$`\frac{1}{4}\delta _{ij}\left(2g^2[2T_{nk}T_{nk}(T_{kk})^2]ฯต_{D3}T_\mathrm{}m^1T_{nk}^1F_{\left(2\right)}^\mathrm{}kF_{\left(2\right)}^{mn}\right),`$
$`d(Y^1F_{\left(3\right)})=0.`$ (22)
We have introduced the unimodular matrix $`\stackrel{~}{T}_{ij}`$, constructed from $`T_{ij}`$ by extracting the determinant factor $`Y`$ (see (16)),
$$\stackrel{~}{T}_{ij}=Y^{\frac{1}{4}}T_{ij}.$$
(23)
Again, there are many other terms that arise from acting with the exterior derivative that cancel amongst themselves, after making use of the Schoutens identity. The consistency of the reduction Ansatz manifests itself in the remarkable fact that one reads off consistent lower-dimensional equations of motion in which all the dependence on the internal $`S^3`$ coordinates $`\mu ^i`$ has cancelled.
Next, we consider the equation of motion for the dilaton $`\widehat{\varphi }`$. From (3) we find
$$d\widehat{\varphi }=\sqrt{\frac{2}{D2}}\left(\frac{1}{4}(D4)Y^1dY\mathrm{\Delta }^1d\mathrm{\Delta }\right).$$
(24)
Since $`\mathrm{\Delta }`$ has dependence on the $`S^3`$ coordinates $`\mu ^i`$ one of the terms arising here will involve the quantity
$$T_{ij}\mu ^i๐\mu ^j.$$
(25)
It is therefore necessary to evaluate the Hodge dual of this 1-form; we find
$$\widehat{}(T_{ij}\mu ^i๐\mu ^j)=\frac{1}{2}ฯต_{i_1i_2i_3i_4}T_i\mathrm{}\mu ^iฯต_{D3}(\mathrm{\Delta }T_{i_1\mathrm{}}T_{i_1j}T_k\mathrm{}\mu ^j\mu ^k)\mu ^{i_2}๐\mu ^{i_3}๐\mu ^{i_4}.$$
(26)
After some involved manipulations, we find that the $`D`$-dimensional dilaton equation of motion in (19) implies that $`Y`$ satisfies
$`\frac{D5}{4}(1)^Dd(Y^1dY)`$ $`=`$ $`\frac{1}{2}g^2(2T_{ij}T_{ij}(T_{ii})^2)ฯต_{D3}Y^1F_{\left(3\right)}F_{\left(3\right)}`$ (27)
$`\frac{1}{4}Y^1T_{ik}^1T_j\mathrm{}^1F_{\left(2\right)}^{ij}F_{\left(2\right)}^k\mathrm{}.`$
The full set of $`(D3)`$-dimensional equations of motion can be derived from the Lagrangian
$`_{D3}`$ $`=`$ $`R\text{1}\mathrm{l}\frac{D5}{16}Y^2dYdY\frac{1}{4}\stackrel{~}{T}_{ij}^1๐\stackrel{~}{T}_{jk}\stackrel{~}{T}_k\mathrm{}^1๐\stackrel{~}{T}_\mathrm{}i`$ (28)
$`\frac{1}{2}Y^1F_{\left(3\right)}F_{\left(3\right)}\frac{1}{4}Y^{\frac{1}{2}}\stackrel{~}{T}_{ik}^1\stackrel{~}{T}_j\mathrm{}^1F_{\left(2\right)}^{ij}F_{\left(2\right)}^k\mathrm{}V\text{1}\mathrm{l},`$
where the potential $`V`$ is given by
$$V=\frac{1}{2}g^2Y^{\frac{1}{2}}\left(2\stackrel{~}{T}_{ij}\stackrel{~}{T}_{ij}(\stackrel{~}{T}_{ii})^2\right).$$
(29)
The 3-form field strength $`F_3`$ is given by
$$F_{\left(3\right)}=dA_{\left(2\right)}+\frac{1}{8}ฯต_{ijk\mathrm{}}(F_{\left(2\right)}^{ij}A_{\left(1\right)}^k\mathrm{}\frac{1}{3}gA_{\left(1\right)}^{ij}A_{\left(1\right)}^{km}A_{\left(1\right)}^m\mathrm{}),$$
(30)
which implies that $`F_{\left(3\right)}`$ satisfies the Bianchi identity (20).
## 4 Truncations to previous results
In this section, we consider two truncations of the $`S^3`$ Kaluza-Klein reduction of the bosonic string that we constructed in the previous section.
### 4.1 Truncation from $`SO(4)`$ to $`SU(2)`$
The first truncation turns the reduction into a โstandardโ one, for which the consistency becomes immediately understandable from group-theoretic arguments. Specifically, we may truncate the $`SO(4)`$ Yang-Mills gauge fields that arise from the $`S^3`$ reduction to a set of $`SU(2)`$ gauge fields, corresponding either to the left-action, or to the right-action, of $`SU(2)`$ on the $`S^3SU(2)`$ group manifold. This is achieved by imposing a self-dual or anti-self-dual truncation on the original $`SO(4)`$ gauge potentials $`A_{\left(1\right)}^{ij}`$,
$$A_{\left(1\right)}^{ij}=\pm \frac{1}{2}ฯต_{ijk\mathrm{}}A_{\left(1\right)}^k\mathrm{}.$$
(31)
The choice of sign governs whether we are retaining the gauge fields of $`SU(2)_L`$ or of $`SU(2)_R`$ in the truncation of $`SO(4)SU(2)_L\times SU(2)_R`$. The two choices are equivalent, up to convention choices, and we shall pick the plus sign in (31) for definiteness. It is convenient to take the $`i,j,\mathrm{}`$ indices to range over the values $`0,1,2,3`$, and to write the remaining gauge potentials in terms of the $`SU(2)`$ triplet $`A_{\left(1\right)}^\alpha `$, with
$$A_{\left(1\right)}^{01}=A_{\left(1\right)}^{23}\frac{1}{2}A_{\left(1\right)}^1,A_{\left(1\right)}^{02}=A_{\left(1\right)}^{31}\frac{1}{2}A_{\left(1\right)}^2,A_{\left(1\right)}^{03}=A_{\left(1\right)}^{12}\frac{1}{2}A_{\left(1\right)}^3.$$
(32)
These are the gauge fields of $`SU(2)_L`$.
At the same time as we impose the self-dual truncation (31) on the gauge potentials, we must also truncate the scalar fields $`T_{ij}`$, in order to be consistent with the equations of motion for the truncated gauge fields. In fact we should retain just a single scalar degree of freedom $`X`$, so that $`T_{ij}`$ becomes
$$T_{ij}=X\delta _{ij}.$$
(33)
Note that from (16) we shall now have $`Y=X^4`$. It is convenient also to give an explicit parametrisation of the $`\mu ^i`$ coordinates, in terms of Euler angles on $`S^3`$:
$$\mu _0+\mathrm{i}\mu _3=\mathrm{cos}\frac{1}{2}\theta e^{\mathrm{i}(\psi +\varphi )/2},\mu _1+\mathrm{i}\mu _2=\mathrm{sin}\frac{1}{2}\theta e^{\mathrm{i}(\psi \varphi )/2}.$$
(34)
In terms of these we can then define the left-invariant 1-forms $`\sigma _\alpha `$ on $`S^3`$, according to
$$\sigma _1+\mathrm{i}\sigma _2=e^{\mathrm{i}\psi }(d\theta +\mathrm{i}\mathrm{sin}\theta d\varphi ),\sigma _3=d\psi +\mathrm{cos}\theta d\varphi .$$
(35)
These satisfy the $`SU(2)`$ algebra $`d\sigma _\alpha =\frac{1}{2}ฯต_{\alpha \beta \gamma }\sigma _\beta \sigma _\gamma `$.
With these preliminaries, we can now present our results for the reduction Anstaz for this $`SU(2)`$ truncation of the original $`SO(4)`$ Kaluza-Klein reduction. We find that the metric, dilaton and 3-form Ansรคtze given in (3)-(15) reduce to
$`d\widehat{s}_D^2`$ $`=`$ $`X^{\frac{6}{D2}}ds_{D3}^2+\frac{1}{4}X^{\frac{2(D5)}{D2}}{\displaystyle \underset{\alpha }{}}(\sigma _\alpha gA_{\left(1\right)}^\alpha )^2,`$ (36)
$`e^{\sqrt{(D2)/2}\widehat{\varphi }}`$ $`=`$ $`X^{D5},`$ (37)
$`\widehat{F}_{\left(3\right)}`$ $`=`$ $`F_{\left(3\right)}{\displaystyle \frac{1}{4g^2}}\mathrm{\Omega }_{\left(3\right)}{\displaystyle \frac{1}{12g}}ฯต_{\alpha \beta \gamma }F_{\left(2\right)}^\alpha (\sigma _\beta gA_{\left(1\right)}^\beta )(\sigma _\gamma gA_{\left(1\right)}^\gamma ),`$ (38)
where
$$\mathrm{\Omega }_{\left(3\right)}\frac{1}{6}ฯต_{\alpha \beta \gamma }(\sigma _\alpha gA_{\left(1\right)}^\alpha )(\sigma _\beta gA_{\left(1\right)}^\beta )(\sigma _\gamma gA_{\left(1\right)}^\gamma )$$
(39)
is the volume form on the 3-sphere.
It is easy to verify that this $`SU(2)`$ truncation of the full $`SO(4)`$ reduction Ansatz of section 2 is a consistent one. As we remarked above, there is no longer anything โsurprisingโ about the consistency in this case, since the truncation has set to zero all fields that transformed non-trivially under $`SU(2)_R`$. In other words, the $`SU(2)_L`$ Ansatz in this section retains all the singlets under $`SU(2)_R`$, while discarding all the non-singlets. Such a truncation is necessarily consistent, since non-linear products of the fields that are retained can never generate non-singlets under $`SU(2)_R`$. A related point is that the fields that remain in the reduction Ansatz parameterise homogeneous deformations of the 3-sphere. A particular case of this $`SU(2)`$ reduction has appeared previously in the literature, in the $`S^3`$ reduction of $`N=1`$ supergravity from $`D=10`$ to $`D=7`$ .
### 4.2 Truncation from $`SO(4)`$ to $`U(1)\times U(1)`$
The second truncation that we shall consider here involves retaining only the $`U(1)\times U(1)`$ subgroup of the original $`SU(2)\times SU(2)`$ gauged fields of the full $`SO(4)`$ reduction Ansatz of section 2. It is convenient now to take the $`SO(4)`$ indices $`i,j,\mathrm{}`$ to range over the values $`1,2,3,4`$. The truncation amounts to setting all gauge potentials $`A_{\left(1\right)}^{ij}`$ to zero except for $`A_{\left(1\right)}^{12}`$ and $`A_{\left(1\right)}^{34}`$, for which we write
$$A_{\left(1\right)}^{12}A_{\left(1\right)}^1,A_{\left(1\right)}^{34}A_{\left(1\right)}^2.$$
(40)
It is also convenient now to parameterise the coordinates $`\mu ^i`$ on $`S^3`$ as
$$\mu _1+\mathrm{i}\mu _2=\stackrel{~}{\mu }_1e^{\mathrm{i}\varphi _1},\mu _3+\mathrm{i}\mu _4=\stackrel{~}{\mu }_2e^{\mathrm{i}\varphi _2}.$$
(41)
At the same time as making the truncation of the gauge fields, consistency with their equations of motion requires that we set certain of the scalar fields to zero, so that what remains is just two scalars $`X_1`$ and $`X_2`$ as follows:
$$T_{ij}=\mathrm{diag}(X_1,X_1,X_2,X_2).$$
(42)
Note that we shall now have
$$Y=(X_1X_2)^2,\mathrm{\Delta }=X_1\stackrel{~}{\mu }_1^2+X_2\stackrel{~}{\mu }_2^2,U=2\underset{i=1}{\overset{2}{}}(X_i^2\stackrel{~}{\mu }_i^2\mathrm{\Delta }X_i).$$
(43)
After substituting the above truncation and reparametrisation into the original Kaluza-Klein Ansรคtze in section 2, we find that the metric and dilaton Asรคtze become
$`d\widehat{s}_D^2`$ $`=`$ $`(X_1X_2)^{\frac{2}{D2}}(\mathrm{\Delta }^{\frac{2}{D2}}ds_{D3}^2`$ (44)
$`+g^2\mathrm{\Delta }^{\frac{D4}{D2}}{\displaystyle \underset{i=1}{\overset{2}{}}}X_i^1(d\stackrel{~}{\mu }_i^2+\stackrel{~}{\mu }_i^2(d\varphi _igA_{\left(1\right)}^i)^2)),`$
$`e^{\sqrt{(D2)/2}\widehat{\varphi }}`$ $`=`$ $`\mathrm{\Delta }^1(X_1X_2)^{\frac{D4}{2}}.`$ (45)
The Ansatz for the 3-form field $`\widehat{F}_{\left(3\right)}`$ in this $`U(1)^2`$ truncation is most simply expressed in terms of the expression for the dual of $`\widehat{F}_{\left(3\right)}`$. Making the truncation in (3), we find
$`e^{\sqrt{8/(D2)}\widehat{\varphi }}\widehat{}\widehat{F}_{\left(3\right)}`$ $`=`$ $`2g{\displaystyle \underset{i=1}{\overset{2}{}}}\left(X_i^2\stackrel{~}{\mu }_i^2\mathrm{\Delta }X_i\right)ฯต_{D3}+{\displaystyle \frac{1}{2g}}{\displaystyle \underset{i=1}{\overset{2}{}}}X_i^1dX_id(\stackrel{~}{\mu }_i^2)`$ (46)
$`{\displaystyle \frac{1}{2g^2}}{\displaystyle \underset{i=1}{\overset{2}{}}}X_i^2d(\stackrel{~}{\mu }_i^2)(d\varphi _igA_{\left(1\right)}^i)F_{\left(2\right)}^i+g^3Y^1F_{\left(3\right)}.`$
The Ansatz for $`\widehat{F}_{\left(3\right)}`$ itself is also easily obtainable by imposing the $`U(1)^2`$ truncation on the general $`SO(4)`$ Ansatz (15).
Note that in the $`U(1)^2`$ truncation the question of the consistency of the reduction is still a non-trivial one, since the two scalars $`X_1`$ and $`X_2`$ parameterise inhomogeneous deformations of the 3-sphere. Of course since we have already argued that the $`SO(4)`$ reduction in section 2 is consistent, the consistency for the $`U(1)^2`$ truncation is a consequence.
A particular case of this $`U(1)^2`$ truncation appeared previously in the literature , where it was obtained for the case $`D=10`$ by taking a singular limit of the $`S^4`$ reduction of eleven-dimensional supergravity that was constructed in .
## 5 $`S^{D3}`$ reduction and $`D=3`$, $`N=8`$ gauged supergravity
As we discussed in section 2, it is natural to conjecture that the theory of gravity coupled to a dilaton and a 3-form, described by (14) with $`a`$ given by (7), should also admit a consistent reduction to three dimensions on the sphere $`S^{D3}`$, in which all the Yang-Mills gauge fields of $`SO(D2)`$ are retained. Additionally, we should keep the $`\frac{1}{2}(D1)(D2)`$ scalar fields described by the symmetric tensor $`T_{ij}`$, where $`i`$ is a vector index of $`SO(D2)`$. We find that indeed such a consistent reduction is possible, and that the Kaluza-Klein Ansatz is given by
$`d\widehat{s}_D^2`$ $`=`$ $`Y^{\frac{1}{D2}}\left(\mathrm{\Delta }^{\frac{D4}{D2}}ds_3^2+g^2\mathrm{\Delta }^{\frac{2}{D2}}T_{ij}^1๐\mu ^i๐\mu ^j\right),`$
$`e^{\sqrt{(D2)/2}\widehat{\varphi }}`$ $`=`$ $`\mathrm{\Delta }^1Y^{1/2},`$
$`\widehat{F}_{\left(3\right)}`$ $`=`$ $`gUฯต_3+g^1T_{ij}^1๐T_{jk}(\mu ^k๐\mu ^i)`$ (47)
$`{\displaystyle \frac{1}{2g^2}}T_{ik}^1T_{jl}^1F_{\left(2\right)}^{ij}๐\mu ^k๐\mu ^{\mathrm{}},`$
where the various quantities appearing here are again given in (16), (17) and (18), but now the indices $`i,j,\mathrm{}`$ range over $`(D2)`$ values. Thus the field content in Kaluza-Klein reduced three-dimensional theory comprises the metric $`ds_3^2`$, the $`\frac{1}{2}(D2)(D3)`$ gauge potentials $`A_{\left(1\right)}^{ij}`$ of $`SO(D2)`$, and the $`\frac{1}{2}(D1)(D2)`$ scalars described by the symmetric tensor $`T_{ij}`$. The calculation of the Hodge dual of the 3-form $`\widehat{F}_{\left(3\right)}`$ is again a mechanical, although involved, calculation. We find that it is given by
$`e^{\sqrt{8/(D2)}\widehat{\varphi }}\widehat{}\widehat{F}_{\left(3\right)}`$ $`=`$ $`{\displaystyle \frac{g^{(D4)}}{(D3)!}}ฯต_{i_1\mathrm{}i_{D2}}(gU\mathrm{\Delta }^2\mu ^{i_1}๐\mu ^{i_2}\mathrm{}๐\mu ^{i_{D2}}`$ (48)
$`\left(D3\right)\mathrm{\Delta }^2T_{i_1j}๐T_{i_2k}๐\mu ^{i_3}\mathrm{}๐\mu ^{i_{D2}}\mu ^j\mu ^k`$
$`\frac{(D3)(D4)}{2}F_{\left(2\right)}^{i_1i_2}T_{i_3j}๐\mu ^{i_4}๐\mu ^{i_{D2}}\mu ^j),`$
where we have suppressed the wedge symbols in products of differential forms in order to economise on space.
It is again a straightforward, although lengthy, procedure to substitute the above Ansatz into the $`D`$-dimensional equations of motion (19), and to verify that there is a consistent reduction to equations of motion for the three-dimensional fields. We find that these equations can be derived from the following three-dimensional Lagrangian:
$`_3`$ $`=`$ $`R\text{1}\mathrm{l}\frac{1}{4(D2)}Y^2dYdY\frac{1}{4}\stackrel{~}{T}_{ij}^1๐\stackrel{~}{T}_{jk}\stackrel{~}{T}_k\mathrm{}^1๐\stackrel{~}{T}_\mathrm{}i`$ (49)
$`\frac{1}{4}Y^{\frac{2}{D2}}\stackrel{~}{T}_{ik}^1\stackrel{~}{T}_j\mathrm{}^1F_{\left(2\right)}^{ij}F_{\left(2\right)}^k\mathrm{}V\text{1}\mathrm{l},`$
where $`Y=det(T_{ij})`$, and $`T_{ij}`$ is written in terms of the unimodular $`(D2)\times (D2)`$ matrix $`\stackrel{~}{T}_{ij}`$ as $`T_{ij}=Y^{1/(D2)}\stackrel{~}{T}_{ij}`$. The potential $`V`$ is given by
$$V=\frac{1}{2}g^2Y^{\frac{2}{D2}}\left(2\stackrel{~}{T}_{ij}\stackrel{~}{T}_{ij}(\stackrel{~}{T}_{ii})^2\right).$$
(50)
An application of this dimensional reduction that is of particular interest arises if we take $`D=10`$, since then the higher-dimensional starting point will be the bosonic sector of $`N=1`$ supergravity in ten dimensions. The reduction on $`S^7`$ then yields a three-dimensional theory that is the bosonic sector of an $`SO(8)`$ gauged supergravity, with $`N=8`$ (i.e. half of maximal) supersymmetry. As well as the 28 gauge fields, there are in total 36 scalars, described by the unimodular symmetric tensor $`\stackrel{~}{T}_{ij}`$ and the scalar $`Y`$. These transform as a 35 and a 1 under $`SO(8)`$, respectively. Evidently, if we reduced the full $`N=1`$ theory in $`D=10`$, including the fermions, we would obtain $`N=8`$ gauged $`SO(8)`$ supergravity in three dimensions. This appears to be the first example of such a gauged supergravity in $`D=3`$. Previous examples of gauged three-dimensional supergravities in the literature have been of the type constructed in , with $`SO(p)\times SO(q)`$ gauge fields and a pure cosmological constant term implying the existence of an AdS<sub>3</sub> ground-state solution. In fact there are no scalar fields, and hence no scalar potential, in the theories constructed in . By contrast, the gauged supergravity that we have obtained here has 36 scalars with the potential (50). The theory does not admit an AdS<sub>3</sub> solution, but it may allow domain-wall solutions that preserve half of the supersymmetry.
## 6 $`S^2`$ reduction
Here, we construct the Kaluza-Klein Ansatz for the reduction of (3) with $`p=2`$ and $`a`$ given by (6). Thus our starting point is
$$_D=\widehat{R}\widehat{}\text{1}\mathrm{l}\frac{1}{2}\widehat{}d\widehat{\varphi }d\widehat{\varphi }\frac{1}{2}e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(2\right)}\widehat{F}_{\left(2\right)},$$
(51)
where the positive constant $`a`$ is given by (6). From (51) we derive the equations of motion
$`d\widehat{}d\widehat{\varphi }`$ $`=`$ $`\frac{1}{2}(1)^Dae^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(2\right)}\widehat{F}_{\left(2\right)},`$
$`d(e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(2\right)})`$ $`=`$ $`0,`$ (52)
$`\widehat{R}_{MN}`$ $`=`$ $`\frac{1}{2}_M\widehat{\varphi }_N\widehat{\varphi }+\frac{1}{2}e^{a\widehat{\varphi }}\left(\widehat{F}_{MN}^1\frac{1}{2(D2)}\widehat{F}_{\left(2\right)}^2\widehat{g}_{MN}\right).`$ (53)
We find that there is a consistent reduction Ansatz on $`S^2`$, given by
$`d\widehat{s}_D^2`$ $`=`$ $`Y^{\frac{1}{D2}}\left(\mathrm{\Delta }^{\frac{1}{D2}}ds_{D2}^2+g^2\mathrm{\Delta }^{\frac{D3}{D2}}T_{ij}^1๐\mu ^i๐\mu ^j\right),`$ (54)
$`e^{\sqrt{\frac{2(D2)}{D1}}\widehat{\varphi }}`$ $`=`$ $`\mathrm{\Delta }^1Y^{\frac{D3}{D1}},`$ (55)
$`\widehat{F}_{\left(2\right)}`$ $`=`$ $`\frac{1}{2}ฯต_{ijk}(g^1U\mathrm{\Delta }^2\mu ^i๐\mu ^j๐\mu ^k2g^1\mathrm{\Delta }^2๐\mu ^i๐T_j\mathrm{}T_{km}\mu ^{\mathrm{}}\mu ^m`$ (56)
$`\mathrm{\Delta }^1F_{\left(2\right)}^{ij}T_k\mathrm{}\mu ^{\mathrm{}}).`$
Again, the various quantities appearing here are given in (16), (17) and (18), but with the indices $`i,j,\mathrm{}`$ ranging over 3 values. The dual of the 2-form then turns out to be given by
$`e^{\sqrt{\frac{2(D1)}{D2}}\widehat{\varphi }}\widehat{}\widehat{F}_{\left(2\right)}`$ $`=`$ $`gUฯต_{D2}+g^1T_{ij}^1๐T_{jk}(\mu ^k๐\mu ^i)`$ (57)
$`{\displaystyle \frac{1}{2g^2}}T_{ik}^1T_{jl}^1F_{\left(2\right)}^{ij}๐\mu ^k๐\mu ^{\mathrm{}}.`$
The field content of the Kaluza-Klein reduced theory comprises the $`(D2)`$-dimensional metric $`ds_{D2}^2`$, the three gauge potentials $`A_{\left(1\right)}^{ij}`$ of $`SO(3)`$, and the six scalar fields $`T_{ij}`$.
Substituting the Ansatz into the $`D`$-dimensional equations of motion (52), we find that it yields a consistent Kaluza-Klein $`S^2`$ reduction, with the $`(D2)`$-dimensional fields satisfying equations of motion that follow from the Lagrangian
$`_{D2}`$ $`=`$ $`R\text{1}\mathrm{l}\frac{D4}{3(D1)}Y^2dYdY\frac{1}{4}\stackrel{~}{T}_{ij}^1๐\stackrel{~}{T}_{jk}\stackrel{~}{T}_k\mathrm{}^1๐\stackrel{~}{T}_\mathrm{}i`$ (58)
$`\frac{1}{4}Y^{\frac{2}{3}}\stackrel{~}{T}_{ik}^1\stackrel{~}{T}_j\mathrm{}^1F_{\left(2\right)}^{ij}F_{\left(2\right)}^k\mathrm{}V\text{1}\mathrm{l},`$
where $`Y=det(T_{ij})`$, and $`T_{ij}`$ is written in terms of the unimodular $`3\times 3`$ matrix $`\stackrel{~}{T}_{ij}`$ as $`T_{ij}=Y^{1/3}\stackrel{~}{T}_{ij}`$. The potential $`V`$ is given by
$$V=\frac{1}{2}g^2Y^{\frac{2}{3}}\left(2\stackrel{~}{T}_{ij}\stackrel{~}{T}_{ij}(\stackrel{~}{T}_{ii})^2\right).$$
(59)
In view of our earlier observation that the $`D`$-dimensional Lagrangian (51), with the constant $`a`$ given by (6), can itself be thought of as an ordinary $`S^1`$ Kaluza-Klein reduction of pure gravity in $`(D+1)`$ dimensions, it follows that we can also interpret our result as a consistent reduction of $`(D+1)`$-dimensional pure gravity. The internal space is not simply $`S^1\times S^2`$, however, since the 2-form field $`F_{\left(2\right)}`$ in $`D`$ dimensions, which is the Kaluza-Klein vector of the $`S^1`$ reduction from $`(D+1)`$ dimensions, is topologically non-trivial. One can see from (56) that if, for example, the scalars were all taking trivial values, the 2-form field $`\widehat{F}_{\left(2\right)}`$ would be just the volume-form on $`S^2`$ (like in a Dirac monopole configuration). Thus the reduction from $`(D+1)`$ dimensions is actually on a manifold that is topologically $`S^3`$. In fact we can easily lift the metric Ansatz in (54) to give the Ansatz for the reduction from $`(D+1)`$ dimensions, by incorporating the standard $`S^1`$ reduction step
$$d\widehat{s}_{D+1}^2=e^{2\alpha \widehat{\varphi }}d\widehat{s}_D^2+e^{2\alpha (D2)\widehat{\varphi }}(dz+\widehat{A}_{\left(1\right)})^2,$$
(60)
where $`\widehat{F}_{\left(2\right)}=d\widehat{A}_{\left(1\right)}`$, and the fields on the right-hand side are given in (54)โ(56). Thus we find
$$d\widehat{s}_{D+1}^2=Y^{\frac{2}{D1}}ds_{D2}^2+\mathrm{\Delta }^1Y^{\frac{2}{D1}}T_{ij}^1๐\mu ^i๐\mu ^j+\mathrm{\Delta }Y^{\frac{D3}{D1}}(dz+\widehat{A}_{\left(1\right)})^2.$$
(61)
This is an unusual type of $`S^3`$ reduction, in which the three $`SO(3)`$ Yang-Mills fields $`A_{\left(1\right)}^{ij}`$ and the six scalar fields $`T_{ij}`$ parameterise inhomogeneous deformations of the 3-sphere.
## 7 Conformal anomaly terms
Until now we have focussed our attention on the purely classical theories of gravity coupled to a $`p`$-form field strength and a dilaton. One of the two cases that admits consistent sphere reductions turned out to be when $`p=3`$, and in fact the Lagrangian (14) is precisely the leading-order expression for the low-energy limit of the $`D`$-dimensional bosonic string. Of course the bosonic string suffers from a conformal anomaly if the dimension $`D`$ is not equal to 26. It turns out that the effect of this anomaly is to generate an additional term in the effective action , which vanishes at $`D=26`$, so that (14) is replaced by
$$_D=\widehat{R}\widehat{}\text{1}\mathrm{l}\frac{1}{2}\widehat{}d\widehat{\varphi }d\widehat{\varphi }\frac{1}{2}e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(3\right)}\widehat{F}_{\left(3\right)}\frac{1}{2}m^2(D26)e^{\frac{1}{2}a\widehat{\varphi }}\widehat{}\text{1}\mathrm{l}.$$
(62)
We shall refer to this extra contribution as a โcosmological term.โ Note that if we were instead considering the theory of gravity, 3-form and dilaton as coming from the low-energy effective theory of the superstring, the $`(D26)`$ factor would be replaced by $`(D10)`$. In all subsequent discussions in this section, 26 can accordingly be replaced by 10 in the context of the superstring.
It is of interest to see what happens to the previous Kaluza-Klein reductions on $`S^3`$ and $`S^{D3}`$ after this extra term is included. We find that the previous $`S^3`$ reduction Ansatz continues to give a consistent reduction, in which all the dependence on the $`S^3`$ coordinates cancels out when it is substituted into the $`D`$-dimensional equations of motion following from (62). We find that the reduced $`(D3)`$-dimensional theory is described by the same Lagrangian (28), but now the potential $`V`$ given in (29) is replaced by
$$V=\frac{1}{2}g^2Y^{\frac{1}{2}}\left(2\stackrel{~}{T}_{ij}\stackrel{~}{T}_{ij}(\stackrel{~}{T}_{ii})^2\right)+\frac{1}{2}m^2(D26)Y^{\frac{1}{2}}.$$
(63)
The fact that the $`S^3`$ reduction continues to be a consistent one after the inclusion of the cosmological term in (62) could in fact have been foreseen by considering the group-theory arguments that we developed in section 2. In the absence of the cosmological term, we observed that the global symmetry group after a $`T^3`$ reduction is $`\text{I}\mathrm{R}\times O(3,3)`$, which is large enough to contain $`O(3)\times O(3)`$ as a compact subgroup, and hence to permit an $`SO(4)`$ gauging. The inclusion of the cosmological term in (62) breaks the $`\text{I}\mathrm{R}`$ factor in the global symmetry, but the $`O(3,3)`$ factor survives,<sup>11</sup><sup>11</sup>11This can be seen from the fact that the dilaton vector for the cosmological term after the $`T^3`$ reduction is orthogonal to the dilaton vectors that form the positive roots of $`O(3,3)`$. and so the cosmological term does not present any obstacle to the $`SO(4)`$ gauging in $`D3`$ dimensions.
It is interesting to note that if $`D>26`$ (or $`D>10`$ in the case of a supersymmetric string), the potential (63) admits a symmetrical ground-state solution in which all the scalar fields are constant. To see this, we note that for such a solution we must have
$$\frac{V}{Y}=0,\frac{V}{\stackrel{~}{T}_{ij}}\frac{1}{4}\delta _{ij}\delta _k\mathrm{}\frac{V}{\stackrel{~}{T}_k\mathrm{}}=0.$$
(64)
(The trace subtraction in the second equation arises because $`\stackrel{~}{T}_{ij}`$ has unit determinant.) Thus the conditions for a solution with constant scalars imply
$$V=0,\stackrel{~}{T}_{ij}=\frac{1}{4}\stackrel{~}{T}_{kk}\delta _{ij},$$
(65)
and hence since $`\stackrel{~}{T}_{ij}`$ is unimodular we must have $`\stackrel{~}{T}_{ij}=\delta _{ij}`$, and
$$g=m\sqrt{\frac{D26}{8}},$$
(66)
with $`Y`$ arbitrary. Note in particular that the vanishing of $`V`$ implies that the $`(D3)`$-dimensional Einstein equation has no cosmological term, and so it admits Minkowski spacetime as a ground-state solution. One can also find non-trivial solutions that are asymptotically flat.
If we now consider instead the $`S^{D3}`$ reduction of the new theory (62), we find that the previously consistent reduction is spoiled by the presence of the additional cosmological term. In particular, it turns out that there is a mis-match between the $`S^{D3}`$ dependence from the extra term $`e^{\frac{1}{2}a\widehat{\varphi }}\widehat{}\text{1}\mathrm{l}`$, in comparison to the previous terms, in the $`D`$-dimensional equation of motion for the dilaton $`\widehat{\varphi }`$. Actually, this is not too surprising. It can be understood from the fact that the presence of the cosmological term in (62) breaks the enhanced $`O(D2,D2)`$ global symmetry that occurred previously under a dimensional reduction on $`T^{D3}`$, and so there will no longer be an $`SO(D2)`$ compact subgroup of the global symmetry group that could permit an $`SO(D2)`$ gauging in three dimensions. This can be seen from the fact that the dilaton vector for the cosmological term in (62), after toroidal reduction on $`T^{D3}`$, is not orthogonal to the positive root vectors of $`O(D2,D2)`$.
Finally, we may also consider the possible inclusion of an analogous cosmological term in the Lagrangian (51) for gravity, the dilaton and a 2-form field strength. In this case there would not be any direct motivation from bosonic string theory for the inclusion of such a term, but it is nevertheless of interest to see what the effect would be. Thus we may consider whether we may modify the Lagrangian (51) to
$$_D=\widehat{R}\widehat{}\text{1}\mathrm{l}\frac{1}{2}\widehat{}d\widehat{\varphi }d\widehat{\varphi }\frac{1}{2}e^{a\widehat{\varphi }}\widehat{}\widehat{F}_{\left(2\right)}\widehat{F}_{\left(2\right)}\frac{1}{2}m^2e^{b\widehat{\varphi }}\widehat{}\text{1}\mathrm{l},$$
(67)
where the dilaton coupling constant $`b`$ in the cosmological term is chosen so as to maintain the consistency of the Kaluza-Klein $`S^2`$ reduction. It turns out that this is indeed possible, and consistency is achieved if $`b`$ is the positive constant given by
$$b^2=\frac{2}{(D1)(D2)}.$$
(68)
The resulting Kaluza-Klein theory in $`(D2)`$ dimensions is described by the Lagrangian (58), but with the potential (59) replaced by
$$V=\frac{1}{2}g^2Y^{\frac{2}{3}}\left(2\stackrel{~}{T}_{ij}\stackrel{~}{T}_{ij}(\stackrel{~}{T}_{ii})^2\right)+\frac{1}{2}m^2Y^{\frac{2}{D2}}.$$
(69)
Again, one could have foreseen the continued consistency of the $`S^2`$ reduction from the fact that if the theory (67) is reduced instead on $`T^2`$, there is still a sufficient enhancement of the global symmetry to permit an $`SO(3)`$ gauging. Previously, for (51), the generic $`GL(2,\text{I}\mathrm{R})`$ symmetry was enhanced to to $`GL(3,\text{I}\mathrm{R})`$. Now, with the inclusion of the cosmological term in (67), the $`\text{I}\mathrm{R}`$ factor in the $`GL(3,\text{I}\mathrm{R})`$ is broken, but the $`SL(3,\text{I}\mathrm{R})`$ factor remains, and so the compact $`SO(3)`$ subgroup is still available for the gauging. We can also understand this as follows. Recalling that the original Lagrangian (51) can itself be viewed as a standard $`S^1`$ reduction of pure gravity in $`(D+1)`$ dimensions, we now observe that the enlarged Lagrangian (67), with $`b`$ given by (68), is nothing but the $`S^1`$ reduction of the $`(D+1)`$-dimensional theory of pure gravity with a cosmological constant:
$$_{D+1}=\widehat{R}_{D+1}\widehat{}\text{1}\mathrm{l}\frac{1}{2}m^2\widehat{}\text{1}\mathrm{l}.$$
(70)
It is then evident that the dimensional reduction of (67) on $`T^2`$ will give the same theory as the dimensional reduction of (70) on $`T^3`$, and so in particular there will be a $`SL(3,\text{I}\mathrm{R})`$ global symmetry.<sup>12</sup><sup>12</sup>12The cosmological constant in $`(D+1)`$ dimensions breaks the scale-covariance that a theory of gravity and antisymmetric tensors has, and so one only gets $`SL(n,\text{I}\mathrm{R})`$, and not $`GL(n,\text{I}\mathrm{R})`$ from a $`T^n`$ reduction in this case (see ).
One can again look for solutions of the reduced theory in which all the scalars are constant. The equations of motion following from (69) then imply that
$$\stackrel{~}{T}_{ij}=\delta _{ij},Y^{\frac{2(D5)}{D2}}=\frac{m^2}{g^2(D2)}.$$
(71)
Substituting these back into the potential, we find that at this extremum it is give by
$$V=\frac{1}{2}m^2\left(\frac{D5}{D2}\right)\left[\frac{m^2}{g^2(D5)}\right]^{\frac{1}{D5}},$$
(72)
which corresponds (for $`D6`$) to a positive cosmological constant in the $`(D2)`$-dimensional spacetime. (Note that (71) implies that the cosmological constant in the $`(D+1)`$-dimensional pure gravity theory is also positive.) This allows, in particular, a ground-state solution of the original $`D`$-dimensional theory of the form $`M_{D2}\times S^2`$, where $`M_{D2}`$ is an Einstein spacetime with positive cosmological constant, such as de Sitter space. Interpreted as a solution of the $`(D+1)`$-dimensional pure Einstein theory with positive cosmological constant, it becomes $`M_{D2}\times S^3`$, since in this solution the 2-form $`\widehat{F}_{\left(2\right)}`$ in $`D`$ dimensions is a constant multiple of the volume-form of $`S^2`$, and thus the $`S^1`$ in the reduction from $`(D+1)`$ dimensions is the Hopf bundle over $`S^2`$.
## 8 Conclusions and discussions
In this paper, we have investigated the consistency of the Kaluza-Klein sphere reduction of the theory described by the Lagrangian (3), comprising gravity coupled to a $`p`$-form field strength and a dilaton in $`D`$ dimensions. Specifically, we have focussed our attention on those cases where the reduction Ansatz at least includes all the Yang-Mills fields of the $`SO(n+1)`$ gauge group.
We have shown that by including the dilaton in the higher-dimensional theory, the possibilities for consistent sphere reductions are extended somewhat, in comparison to the case where the higher-dimensional starting point comprises only gravity and a $`p`$-form field strength. Specifically, if no dilaton is included the only possibilities for consistent sphere reductions of the kind we are considering are those associated with the $`S^4`$ and $`S^7`$ reductions of $`D=11`$ supergravity, and the $`S^5`$ reduction of type IIB supergravity. With the dilaton included, we find that consistent $`S^2`$ reductions are possible for the case of a 2-form in the higher dimension $`D`$, and that consistent $`S^3`$ and $`S^{D3}`$ reductions are possible for the case of a 3-form in the higher-dimension. These reductions are possible starting from an arbitrary dimension $`D`$, provided that the strength of the dilaton coupling to the 2-form or 3-form field strength is chosen appropriately.
The previously-known consistent sphere reductions from $`D=11`$ with a 4-form, and $`D=10`$ with a self-dual 5-form, were associated with supersymmetric higher-dimensional theories. In the examples that we have obtained in this paper, supersymmetry is clearly not in general playing a rรดle, since the higher-dimensional starting point can be a theory of gravity, a dilaton and a 2-form or 3-form in any arbitrary dimension. It is probably more appropriate, therefore, to characterise the theories that admit consistent sphere reductions by the fact that they have the unusual property of giving rise to lower-dimensional theories with certain enhanced global symmetry groups upon toroidal reduction on $`T^n`$. In particular, a necessary condition for a consistent $`n`$-sphere reduction that retains all the Yang-Mills fields of $`SO(n+1)`$ is that the global symmetry $`GL(n,\text{I}\mathrm{R})`$ of a generic theory reduced on $`T^n`$ must be enhanced to a group whose compact subgroup contains $`SO(n+1)`$. These symmetry enhancements occur only in exceptional cases, when scalars coming from the toroidal reduction of metric โconspireโ with scalars coming from the reduction of the $`p`$-form field strength to give an enhanced global symmetry group. It so happens that this same feature of symmetry enhancement is a central feature also in theories such as $`D=11`$ and type IIB supergravity, and their toroidal reductions.
It should be emphasised that the group-theoretic argument that we have been using in order to determine when a consistent sphere reduction may be possible does not, of itself, provide a guarantee of consistency.<sup>13</sup><sup>13</sup>13Unlike the traditional group theory argument that proves conclusively the consistency of a truncation in which all singlets under a symmetry group are retained, and all non-singlets are truncated. Rather, it can be viewed as providing a proof of inconsistency in cases where the necessary enhancement of the global symmetry group in the associated toroidal reduction does not occur. It is rather striking, however, that in all cases where a suitable sufficiently large global symmetry enhancement does occur, we find that a consistent sphere reduction is possible.
An interesting illustration of this point is provided by the reductions that we considered in section 7, where an additional โcosmological termโ was included in the higher-dimensional theory. The argument based on global symmetry enhancement showed that the $`S^{D3}`$ reduction would no longer be consistent, but that the $`S^3`$ and $`S^2`$ reductions still had the possibility of being consistent. And indeed this is just what we found, when we substituted the Ansรคtze into the equations of motion of the higher-dimensional theories with the cosmological terms included.
Although we have argued that supersymmetry is in some sense not of itself the directly crucial ingredient in the question of consistency, it is, nevertheless, worthwhile to consider further the question of supersymmetry and consistent sphere reductions. As well as the examples of the $`S^4`$ and $`S^7`$ reductions from $`D=11`$, and the $`S^5`$ reduction from $`D=10`$, we can now also consider those examples amongst the reductions constructed in this paper that can be associated with supersymmetric theories. Thus, for instance, we can consider the $`S^2`$ reduction of type IIA supergravity, using the R-R 2-form, and the $`S^3`$ and $`S^7`$ reductions of type I or type II supergravity, using the NS-NS (or R-R in the case of type IIB) 3-form.
Constructing the Kaluza-Klein sphere reduction Ansatz for the fermions in a supergravity theory is a notoriously difficult problem, and even when it is attempted the efforts are rarely extended to include the quartic fermion terms. However, we may construct a general argument to show that once a consistent reduction has been constructed in the bosonic sector, the supersymmetry of the higher-dimensional theory will then guarantee that a consistent reduction including the fermions as well must be possible. The argument is as follows. We know that a sphere reduction in which all fields (massive as well as massless) are retained will necessarily be consistent, and it will give rise to a supersymmetric lower-dimensional theory. Furthermore, we know that all the non-linear couplings between the various lower-dimensional fields will be organised, by virtue of the lower-dimensional supersymmetry, into supersymmetrically-covariant couplings of complete supermultiplets. Now, if we demonstrate in the bosonic sector that there is a consistent truncation to the massless sector (i.e. to the bosonic sector of the massless lower-dimensional supermultiplet), then this means that there are no interaction terms in which powers of the massless bosonic fields (i.e. conserved currents built from the massless fields (see )) couple to linear powers of the massive bosons that are being set to zero. But this in turn implies that in the full theory there can be no interaction terms in which supercurrents built from the massless multiplet couple to linear powers of the massive fields. Thus if one shows that it is consistent to make a sphere reduction in which all the bosons of the massless supermultiplet are retained, then this implies that it must be consistent to make a sphere reduction of the supersymmetric theory in which the entire massless supermultiplet is retained.
One can use this argument to show that the $`S^3`$ and $`S^7`$ reductions of $`N=1`$ ten-dimensional supergravity, which are special cases of our more general results in this paper, will be consistent, as a consequence of our results for the bosonic sectors.
## Acknowledgement
C.N.P. is grateful to the University of Pennsylvania for hospitality during the course of this work. We are grateful to Arta Sadrzadeh and Tuan Tran for extensive discussions. |
warning/0003/astro-ph0003010.html | ar5iv | text | # Lithium and rotation on the subgiant branch
## 1 Introduction
This paper is the second in a series about the study of lithium and rotation in evolved stars based on both new high resolution spectroscopic observations and precise rotational velocities obtained with the CORAVEL spectrometer. In Lรจbre et al. (1999, hereafter Paper I) we derived lithium abundances by spectral synthesis analysis and presented the observational data for a large and homogeneous sample of F, G, and K-type Population I subgiant stars. On the basis of this uniform data set, we could confirm the occurrence of the rotational discontinuity near the spectral type F8IV (Gray & Nagar 1985, De Medeiros & Mayor 1989, 1990), and localize a lithium drop-off around the spectral type G2IV. No clear correlation appeared between the lithium abundance and the rotational velocity (see also De Medeiros et al. 1997).
In the present work we investigate the physical processes that underline the lithium and rotational discontinuities along the subgiant branch. For this purpose we first determine the evolutionary status and individual masses of our sample stars by using the HIPPARCOS parallaxes and by comparing the observational Hertzsprung-Russell diagram with evolutionary tracks computed with the Toulouse-Geneva code (ยง2). This allows us to study very precisely the behaviour of the lithium abundance and of the rotational velocity as a function of effective temperature, stellar mass, metallicity and evolutionary stage. The lithium main features are discussed in ยง3 where we compare the observations with theoretical predictions related to the dilution mechanism (Iben 1967a,b). Finally, the connection between the observed rotation behaviour and the magnetic braking due to the deepening of the convective envelope (Gray 1981, Gray & Nagar 1985) is quantified in ยง4. This study, based on a close examination of the stellar parameters, sheds new light on the question of the link between rotation and lithium discontinuities in subgiant stars.
## 2 Observational data and evolutionary status
### 2.1 Working sample
Our analysis is based on the observational data from Paper I. This sample is composed of 120 Pop I subgiant stars with F, G and K spectral types which belong to the โCatalogue of rotational and radial velocities for evolved starsโ (De Medeiros & Mayor 1999), as well as to the Bright Star Catalogue (Hoffleit and Jaschek 1982). We thus use the rotational velocities given in Paper I, as well as the values derived for log $`g`$ , $`A_{\mathrm{Li}}`$ and $`T_{\mathrm{eff}}`$ with their respective errors.
### 2.2 Evolutionary status
In order to interpret accurately the observations, we need to know the mass and the evolutionary stage of the sample stars. We use the HIPPARCOS (ESA 1997) trigonometric parallax measurements to locate precisely our objects in the HR diagram. Among our 120 stars, only one object (HD 144071) has no available Hipparcos parallax and is thus rejected from further analysis. Intrinsic absolute magnitudes M<sub>V</sub> are derived from the parallaxes and the m<sub>V</sub> magnitudes given by Hipparcos. We determine the bolometric corrections $`BC`$ by using the Buser and Kuruczโs relation (1992) between $`BC`$ and V-I (again taken from the Hipparcos Catalogue). Finally, we compute the stellar luminosity and the associated error from the sigma error on the parallax. The uncertainties in luminosity lower than $`\pm 0.1`$ have an influence of $`\pm 0.4`$ in the determination of the masses. We show the results of these determinations in Fig. 1 and in Tab. 1. This table displays the stellar masses and luminosities derived for all objects of our sample. Moreover complementary data used for the present analysis (Vsini, $`T_{\mathrm{eff}}`$ , log $`g`$ , \[Fe/H\] and $`A_{\mathrm{Li}}`$ ) can be found in Papier I (see Tables 1 to 3). The error adopted on T<sub>eff</sub> ($`\pm 200`$K) is typical for this class of stars, as already discussed in Paper I.
We have computed evolutionary tracks with the Toulouse-Geneva code for a range of stellar masses between 1 and 4 M and for different metallicities consistent with the range of our sample stars (see Paper I). However, solar composition being relevant to most of the objects of our sample (about 65%), only tracks computed with \[Fe/H\]=0 will be displayed in the figures. The evolution was followed from the Hayashi fully convective configuration. We used the radiative opacities by Iglesias & Rogers (1996), completed with the atomic and molecular opacities by Alexander & Ferguson (1994). The nuclear reactions are from Caughlan & Fowler (1988) and the screening factors are included according to the prescription by Graboske et al. (1993). No transport processes except for the classical convective mixing (with a value of 1.6 for the mixing length parameter) are taken into account.
### 2.3 Discrimination between dwarfs, subgiants and giants among the sample
In Fig. 1 we compare the observational HR diagram with the evolutionary tracks computed with \[Fe/H\] = 0. The dashed line indicates the evolutionary point where the subgiant branch starts and which corresponds to the hydrogen exhaustion in the stellar central regions (i.e., turnoff point). About 30 stars are located below the turnoff line and therefore appear to be genuine dwarfs, although classified as subgiants in the Bright Star Catalogue. On the other hand, about 15 stars located on the right side of the dotted line have started the ascent of the RGB and are thus considered as giants.
## 3 The Lithium main features
Fig. 2 shows the distribution of $`A_{\mathrm{Li}}`$ on the HR diagram and Fig. 3 presents $`A_{\mathrm{Li}}`$ versus log T<sub>eff</sub> for three ranges of metallicity. The main features presented in Paper I, i.e., the lithium discontinuity around log(T<sub>eff</sub>) equal to 3.75, and the dispersion in lithium abundances for subgiants hotter than this value, clearly appear in both figures.
### 3.1 Lithium discontinuity
The observational lithium discontinuity actually simply reflects the well-known dilution that occurs when the convective envelope starts to deepen after the turnoff and reaches the inner free-lithium layers (Iben 1967a,b). In Fig. 4 we show this behaviour in our models as a function of the effective temperature for different masses. A closer look at the models shows that the beginning of the theoretical lithium dilution is a function of stellar mass, as can be seen in Tab. 2 where $`T_{\mathrm{eff}}^{begdil}`$ is the effective temperature at the beginning of the lithium dilution for each mass. We also give $`T_{\mathrm{eff}}^{f10dil}`$, which is the effective temperature at which a decrease of the surface lithium abundance by a factor 10 compared to the value on the main sequence is achieved. The dilution is a very fast process, both in terms of age and effective temperature interval. The theoretical beginning of the lithium dilution is in good agreement with the observed abundance drop-off along the subgiant branch, as can be seen in Fig. 5.
For the considered stellar masses, the predicted dilution factor at the end of the dredge-up (see point a on Fig. 4) ranges between 20 and 60 (these values obtained from our models do not significantly differ from the predictions by Iben 1967a,b). This corresponds to the upper envelope of the observations on the right side of the discontinuity, assuming a cosmic lithium abundance $`A_{\mathrm{Li}}`$ of 3.1.
### 3.2 Lithium dispersion
In the following, we discuss the lithium dispersion on both sides of the discontinuity as a function of the stellar mass inferred from the evolutionary tracks at the corresponding metallicity. We select four ranges of mass according to the ones defined in Balachandran (1995) for the cluster and field stars with respect to the so-called lithium dip region (Boesgaard & Tripicco 1986). Each subsample corresponds to peculiar observational behaviour of the $`A_{\mathrm{Li}}`$ on the main sequence (see Fig. 5 where both main sequence and subgiants stars are shown). In the four mass ranges a large dispersion of the lithium abundance appears and is independent of the single or binary status.
(i) Stars with masses $`<`$ 1.2 M show different degrees of lithium depletion, which occurs already on the pre-main sequence and on main sequence, as observed in open clusters and field stars (Soderblom et al. 1993, Jones et al. 1999 and references therein). This explains their low lithium content when they reach the subgiant branch (see Fig. 5a).
(ii) Stars with masses between 1.2 and 1.5 M correspond to the so-called dip region (Boesgaard & Tripicco 1986).
We separate the stars originating from the hot side of the dip region in two mass ranges :
(iii) stars with masses between 1.5 and 2.25 M, and
(iv) above 2.25 M.
We now discuss in more detail the last three cases.
#### 3.2.1 Li-dip stars
Late F- field and open cluster dwarfs with T<sub>eff</sub> around 6700 K are highly lithium depleted. We find this feature among the dwarfs of our sample which have masses between $``$ 1.2 and 1.5 M (Fig. 5b), in agreement with the observations in the open galactic clusters older than $``$ 200 Myr (Balachandran 1995). This lithium depletion persists in our subgiants of the same mass interval, in agreement with the observations in slightly evolved stars of M67 (Pilachowski et al. 1988, Balachandran 1995, Deliyannis et al. 1997). Explanations relying on the nuclear destruction of lithium are thus favoured by these data (see Talon & Charbonnel 1998 and references therein).
#### 3.2.2 Li in stars originating from the hot side of the dip
As can be seen in Fig. 5c,d (see also Fig. 2), a large lithium dispersion exists among our most massive stars, which show lithium depletion by up to two orders of magnitude before the start of the dilution at Log T$`{}_{\mathrm{eff}}{}^{}3.75`$ (point $`T_{\mathrm{eff}}^{begdil}`$). Very few observational data are available in the literature for stars with masses higher than 1.5 M. In the Hyades, while on the main sequence these objects show lithium abundances close to the galactic value, except for a few deficient Am-stars (Boesgaard 1987, Burkhart & Coupry 1989). However, our observational result is in agreement with the findings by Balachandran (1990) and Burkhart & Coupry (1991) of a few slightly evolved field stars originating from the hot side of the dip and showing significant lithium depletion. We thus confirm the suggestion by Vauclair (1991, see also Charbonnel & Vauclair 1992) that some extra-lithium depletion occurs inside these stars when they are on the main sequence, even if its signature does not appear at the stellar surface at the age of the Hyades. Among some effects suggested one can quote an unusual mass loss on the main sequence (Boesgaard et al. 1977), and rotational induced mixing (Charbonnel & Vauclair 1992, Charbonnel & Talon 1999).
The behaviour observed in our most massive stars which have not yet reached the beginning of dilution explains the very low lithium content of most of the massive subgiants which cannot be accounted for by dilution alone. The underlying destruction mechanism should also be responsible for the lithium behaviour observed in the Hyades giants (Boesgaard et al. 1977) which have masses higher than 2.25 M and thus correspond to the data we show in Fig. 5d. This process should however preserve the boron abundance, which is in agreement with the standard dilution predictions in the underabundant Li giants of the Hyades, as shown recently by Duncan et al. (1998). Observations of boron in our sample stars have to be done to confirm the similarity between field and cluster evolved stars.
Finally, we note that Non-LTE effects alone can not explain the very low Li abundance derived for several massive stars, as proposed by Duncan et al. (1998). Indeed, Carlsson et al. (1994) have showed that Non-LTE effects can reach up to $`+0.3`$ dex only for such cool stars. This correction factor is too low to reconcile the derived Li abundances with the standard dilution predictions. Therefore, this confirms that these stars are indeed Li-poor as well as the hotter ones for which only upper limits have been derived (Non-LTE effects are much smaller for such hotter stars).
## 4 The Vsini main features
In Fig.6 we present the behaviour of rotation in the HR diagram for our sample stars, including the dwarfs and giants. We define several Vsini intervals as in Lรจbre et al. (1999): Slow rotators correspond to Vsini $`<`$ 10 $`\mathrm{km}.\mathrm{s}^1`$, moderate rotators to 10 $`\mathrm{km}.\mathrm{s}^1`$$``$ Vsini $`<`$ 40 $`\mathrm{km}.\mathrm{s}^1`$, and high rotators to Vsini $``$ 40 $`\mathrm{km}.\mathrm{s}^1`$. Fig.6 shows clearly the now well established rotational discontinuity along the subgiant branch near log(T<sub>eff</sub>)=3.79, here indicated by two arrows (Gray & Nagar 1985; De Medeiros & Mayor 1989, 1990). An interesting feature to discuss is the influence of stellar mass on the rotational discontinuity. First of all, one observes that all the subgiant stars with mass lower than about 1.2 M present Vsini values lower than 10.0 $`\mathrm{km}.\mathrm{s}^1`$. Observations in young galactic clusters (see Gaigรฉ 1993 and references therein) show that these low-mass stars actually acquire a slow rotation early on the main sequence, as explained by the magnetic braking scenario for main sequence stars (Kraft 1967, Schrijver and Pols 1993).
Before the rotational discontinuity, the subgiants with mass larger than 1.2 M present a broad range of Vsini values. Because their very thin surface convective envelopes are not an efficient site for magnetic field generation via a dynamo process, these stars are not expected to experience significant angular momentum loss during their main sequence evolution. This result is again in agreement with the data in open clusters.
As can be seen in Tab. 2 (see also Fig. 4), the effective temperature at which the convective envelope starts to deepen (T$`{}_{\mathrm{eff}}{}^{}{}_{}{}^{dee}`$) depends slightly on the stellar mass. We also give the depth of the convective envelope at the effective temperature of the rotational discontinuity, M$`{}_{\mathrm{cz}}{}^{}{}_{}{}^{Rot}`$. For the masses lower than 2.0 M, the observed rotation discontinuity occurs just when the convective envelope starts deepening. We thus see that if the magnetic braking plays a relevant role in the rotational discontinuity, it requires only a very small change in the mass of the convective envelope. Above 2.0 M, our sample is very sparse (due in particular to the very rapid evolution of such stars in the Hertzsprung gap) and we have no data for single stars on the left of the rotation discontinuity exhibited by lower stellar masses. We cannot thus discuss further the impact of the deepening of the convective envelope on the braking in these more massive stars.
## 5 Conclusions
In this work we have analyzed the Li and rotation observations for 120 subgiant stars from Lรจbre et al. (1999). We used the HIPPARCOS trigonometric parallax measurements to locate precisely our objects in the HR diagram and to determine the individual mass and evolutionary status for all stars in the sample.
We have compared observed Li abundances with predictions of Li dilution caused by the deepening of the convective envelope on the subgiant branch. Our models show that the beginning of the theoretical lithium dilution is a function of stellar mass and coincides with the observational features. Stars with masses $`<`$ 1.2 M show a large range in abundance before the turnoff, indicating lithium depletion in the previous phases. Stars with masses between 1.2 and 1.5 M (i.e., in the dip region) show $`A_{\mathrm{Li}}`$ values in agreement with what is found in the open clusters. We note that many stars with masses higher than 1.5 M show lithium depletion up to two orders of magnitude before the start of the dilution at Log T$`{}_{\mathrm{eff}}{}^{}3.75`$. The process that depletes Li in these objects while on the main sequence should however preserve the boron abundance, which is in agreement with the standard dilution predictions in the underabundant Li giants of the Hyades, as shown recently by Duncan et al. (1998). Because these specie burn at different depths than lithium, future observations of boron for our sample will provide powerful additional constraints and will confirm the similarity between field and cluster evolved stars.
Our analysis confirms that low mass stars leave the main sequence with a low rotational rate, while more massive stars are slowed only when reaching the subgiant branch. A very slight increase of the depth of the convective envelope seems to be sufficient for the magnetic braking to take place at this phase. Even if the decrease in the behavior of Lithium is nearly the rotational discontinuity, our interpretation of the observations shows that lithium and rotation discontinuities are independent.
###### Acknowledgements.
J.D.N.Jr. acknowledges partial financial support from the CNPq Brazilian Agency. PdL acknowledges partial support grants from the Sociรฉtรฉ de Secours des Amis des Sciences and the Fondation des Treilles. |
warning/0003/cond-mat0003330.html | ar5iv | text | # The resistance anomaly in the surface layer of Bi2Sr2CaCu2O8+x single crystals under radio-frequency irradiation
## I INTRODUCTION
Recently, the resistance anomaly (RA) characterized by a pronounced resistance enhancement above the normal-state value $`R_N`$ near the superconducting transition has been observed in various systems such as superconducting aluminum wires and structures including artificial normal-metal/superconductor (NS) interfaces. Most of the previous studies on this novel phenomenon indicate that the resistance enhancement takes place either in quasi-one-dimensional structures or in thin films in their superconducting state when the voltage is probed within the range comparable to the characteristic length scales governing the superconducting transition. More recently, it has been found that radio frequency (rf) irradiation controls the appearance and the magnitude of the RA. This leads to explanations for the RA in terms of nonequilibrium superconductivity, such as charge-imbalance phenomenon arising either from rf-excited phase-slip centers or artificial NS interfaces, mixing of extrinsic rf noise, etc.
In this study we have observed similar resistance enhancement in $`c`$-axis conduction measurements on small (junction area of $`25\times 35`$ $`\mu `$m<sup>2</sup>) stacks of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+x</sub> (Bi2212) single crystals irradiated by rf waves, near or below the dc superconducting transition. The RA is observed only for a three-probe measurement configuration, where the voltage measurement includes the surface CuO<sub>2</sub> layer which is in contact with a normal-metal (Au) electrode. The RA in this system resembles the one observed in the conventional aluminum wires in that an excess resistance occurs near or below the superconducting transition and current-voltage ($`IV`$) characteristics exhibit a nonlinear excess voltage at zero bias. Since the anomaly occurs in the $`c`$-axis resistance of a structure including a NS interface (in this case, the interface consists of a Au electrode and the surface CuO<sub>2</sub> layer of Bi2212 single crystals) we believe the RA is related to the charge-imbalance nonequilibrium state produced by rf-induced โhotโ quasiparticles injected from the normal-metal region into the superconducting region. We assume that the rf irradiation excites quasiparticles to higher energy states above the gap of the surface superconducting layer. The enhanced accumulation of charge-imbalanced quasiparticles in the surface CuO<sub>2</sub> layer upon rf irradiation increases the chemical potential of quasiparticles in the layer, which can lead to the enhancement of the measured voltage across the NS interface. The charge-imbalanced nonequilibrium state is known to relax to charge-balanced state within the characteristic time $`\tau _Q^{}=(4/\pi )\tau _i[k_BT/\mathrm{\Delta }(T)]`$ at temperatures near the superconducting transition, where $`\tau _i`$ is the inelastic scattering time, and $`\mathrm{\Delta }(T)`$ the energy gap. The resulting excess resistance due to the charge imbalance becomes most prominent near the superconducting transition, because the relaxation becomes significantly slower as the fraction of the effective states participating in the relaxation, $`\mathrm{\Delta }/k_BT`$, is reduced near the transition.
To our knowledge, no experimental observations of charge-imbalance phenomena of this kind have been reported before in high-$`T_c`$ superconductors (HTSC). This lack of the observation of the phenomena may result from the high phonon-induced inelastic scattering rate near the superconducting transition of HTSC. The electron-phonon scattering rate $`\tau _{eph}^1(\tau _i^1)`$, depending on the quadratic power of the temperature in the two-dimensional clean limit which may be relevant to the surface conduction layer, can be a few hundred times larger in HTSC than in conventional superconductors near the transition. Thus, the probability of HTSC being in a charge-imbalance nonequilibrium state is reduced as much and to date no RA has been reported for the materials. As we reported previously, however, the $`d`$-wave superconductivity of the surface CuO<sub>2</sub> conduction layer in Bi2212 single crystals in contact with a normal-metal electrode is suppressed due to the proximity effect. The superconducting transition temperature of the surface layer, $`T_c^{^{}}`$, is reduced accordingly to at least about one-third of the transition temperature of the inner CuO<sub>2</sub> layers, $`T_c`$. The resultant reduced electron-phonon scattering rate near $`T_c^{^{}}`$ in the surface CuO<sub>2</sub> layer enhances the possibility of observing the nonequilibrium RA effect under rf-wave irradiation.
## II EXPERIMENT
Bi2212 single crystals were grown by the standard solid-state-reaction method. Upon cleaving a Bi2212 single crystal a 50 nm-thick layer of Au was thermally deposited on the top of the crystal to protect the surface from contamination during the further fabrication process as well as to obtain a clean interface between the normal-metal electrodes and the Bi2212 single crystal. Stacks with junction area of $`25\times 35`$ $`\mu `$m<sup>2</sup> and the thickness of about $`15`$ nm were then patterned using the conventional photolithography and Ar-ion-beam etching. A stack of that thickness contains about 10 CuO<sub>2</sub> layers. Further details of the sample fabrication are described elsewhere.
The temperature dependence of the $`c`$-axis resistance $`R_c(T)`$ of a stack was measured by the conventional lock-in technique and $`IV`$ data by a dc method. The schematic measurement configuration is shown in the inset of Fig. 1(a). All measurements were done in a three-terminal configuration with a low pass filter being connected to each measurement electrode. Three-terminal configuration was adopted to probe the surface junction by intentionally including the potential drop across the surface layer. The contact resistance between the Au electrode and the surface CuO<sub>2</sub> layer was less than 0.1 $`\mathrm{\Omega }`$. An rf wave with frequency of 70 MHz was transmitted through coaxial cables to the current leads as shown schematically in the inset of Fig. 1(a). A capacitor was inserted between the specimen and the rf generator to separate the rf signal from the dc probing current. Because of the uncertainty in the rf coupling we were not able to determine the actual rf power transmitted to the specimen. Therefore the power levels specified below are the nominal values.
## III RESULTS AND DISCUSSION
Fig. 1(a) shows the typical $`c`$-axis resistance of a stack of intrinsic junctions, $`R_c(T)`$, in the three-probe configuration with the rf irradiation being turned on (open triangle) or off (open circle). The dc bias current was 500 nA, which corresponds to almost zero bias in the $`IV`$ curves shown in Figs. 2(a) and 2(b). The frequency and the power of the rf wave were 70 MHz and -20 dBm, respectively. In the absence of the rf irradiation it is known that the $`R_c(T)`$ for a stack on the Bi2212 single crystal as illustrated in the inset of Fig. 1(a) exhibits a double superconducting transition; the upper transition at $`T_c`$ ($`90`$ K) is the one for the inner CuO<sub>2</sub> layers and the lower one at $`T_c^{^{}}`$ is for the top-most CuO<sub>2</sub> layer in contact with the Au electrode. The superconductivity in the thin CuO<sub>2</sub> surface layer is suppressed by the proximity contact to the thick normal-metallic Au electrode. Although all the inner layers are Josephson coupled in the temperature range of $`T_c^{^{}}<T<T_c`$, $`R_c(T)`$ remains finite because the surface layer is in the normal state. The value of $`T_c^{^{}}`$ varies from sample to sample. It was 35 K for our specimen in this study.
When an rf wave is irradiated the $`R_c(T)`$ curve deviates below $`T_c^{^{}}`$ from the one in the absence of rf irradiation. Instead of dropping, the $`R_c(T)`$ keeps increasing rapidly below $`T_c^{^{}}`$, before eventually showing a superconducting transition at a temperature further below $`T_c^{^{}}`$. The superconducting transition for a given rf power is much sharper than in the absence of rf irradiation. The $`R_c(T)`$ curve above $`T_c^{^{}}`$ is, however, insensitive to the rf irradiation, which indicates that the excess resistance is caused by the surface CuO<sub>2</sub> layer. As the rf power increases the superconducting transition temperature of the surface CuO<sub>2</sub> layer decreases, along with an increase in the peak value of the RA as shown in Fig. 1(b). For rf power of -20 dBm the RA becomes about 7 times higher than the normal-state resistance at 100 K, which cannot be explained by local heating due to rf irradiation. Since the rf frequency used in this experiment was far lower than the plasma frequency $`\omega _p`$ the rf irradiation may have acted as a dc bias and suppressed the transition temperature. It is also possible that the rf-induced pair breaking across the significantly reduced superconducting gap near $`T_c^{^{}}`$ suppressed the superconducting transition.
We also plotted (dotted line) in Fig. 1(a) the expected $`R_c(T)`$ at temperatures below $`T_c^{^{}}`$ for the normal-metal/insulator/$`d`$-wave superconductor (NID) junction formed by the surface CuO<sub>2</sub> layer in its normal state and the neighboring inner superconducting CuO<sub>2</sub> layer with $`d`$-wave symmetry. A noticeable feature is that the observed resistance enhancement near and below the superconducting transition is far above the estimated tunneling resistance for a NID-junction configuration. The RA, thus, cannot be simply explained by the equilibrium tunneling consideration.
Fig. 2(a) shows the $`IV`$ characteristics at various temperatures below $`T_c^{^{}}`$ when the rf irradiation of frequency 70 MHz and power -20 dBm is on (solid line) or off (dotted line). For 29.8 K data the dotted line is not discernible because of the high overlap of the two sets of curves. Both the $`IV`$ curves at $`T=4.2`$ K and $`T=29.8`$ K show DID and NID tunneling characteristics, respectively, regardless of the rf irradiation. On the other hand, near the superconducting transition (14.9 K - 23.9 K) the $`IV`$ curves in the presence of the rf irradiation show multiple-transition features. For instance, the $`IV`$ curve at 14.9 K shows a double transition, one below and the other above the critical current $`I_c^{^{}}`$ of the intrinsic surface junction in the absence of rf irradiation.
In Figure 2(b) we show a close up view of the zero-bias region of the $`IV`$ characteristics for the three temperatures near the superconducting transition. The $`IV`$ characteristics at $`T=18.9`$ K and $`T=23.9`$ K exhibit excess resistances above the normal-state resistance near the zero bias, which correspond to the RA in the $`R_c(T)`$ near the transition. A resistance enhancement due to the appearance of rf-induced phase-slip centers in only 0.3-nm-thick conducting layer is inconceivable. Burk et al. has proposed that RA can be generated when the high-frequency signal is mixed with the low-frequency or dc measuring current for samples exhibiting nonlinear $`IV`$ characteristics. Since $`IV`$ characteristics of our sample become nonlinear near $`T_c`$ of inner junctions as well as near $`T_c^{^{}}`$ of the surface junction, the model predicts the appearance of the RA near the two temperatures. Since the RA appeared only below $`T_c^{^{}}`$ in our sample, however, we exclude the possibility of external rf mixing, too.
We attribute the RA to the charge imbalance induced in the superconducting surface layer. The excess resistance $`R_{exc}`$ due to the charge imbalance at an interface consisting of a normal metal and a conventional s-wave superconductor is given by
$$R_{exc}=\frac{Z(T)\tau _Q^{}}{2e^2N(0)\mathrm{\Omega }},$$
(1)
where
$$Z(T)=2_\mathrm{\Delta }^{\mathrm{}}N_s(E)\left[\frac{f}{E}\right]๐E,$$
(2)
$`N(0)`$ and $`\mathrm{\Omega }`$ are the quasiparticle density of states of S electrode at the Fermi level in its normal state and the volume of the S electrode, respectively. $`\tau _Q^{}=(4/\pi )\tau _i[k_BT/\mathrm{\Delta }(T)]`$ is the charge-imbalance relaxation time introduced above. It should be noted, however, that Eq. (1) is applicable to a homogeneous superconducting electrode with its length much longer than the charge-imbalance relaxation length $`\mathrm{\Lambda }_Q^{}[=\sqrt{D\tau _Q^{}}50\mu `$m if we assume the diffusivity $`D`$ to be 31 cm<sup>2</sup>/sec\].
We use Eq. (1) to estimate the excess resistance in our specimen, assuming that the charge imbalance is confined in the superconducting surface CuO<sub>2</sub> layer, i.e., $`\mathrm{\Lambda }_Q^{}`$ is set to be the thickness of the surface layer, 0.3 nm. The assumption may be justified by the existence of a strong scattering barrier between the surface and the inner CuO<sub>2</sub> layers. The almost full opening of the superconducting gap in the inner CuO<sub>2</sub> layers in the temperature range near $`T_c^{^{}}`$ also reduces the probability of quasiparticle injection into the inner layers from the surface layer. For the normalized density of states of S, $`N_s(E)`$, we use the angle-averaged expression for the $`d_{x^2y^2}`$ symmetry as $`N_s(E)=Re\left[\frac{1}{2\pi }_0^{2\pi }\frac{E}{\sqrt{E^2[\mathrm{\Delta }(T)\mathrm{cos}(2\theta )]^2}}๐\theta \right]`$ by assuming that the surface CuO<sub>2</sub> layer also has the $`d_{x^2y^2}`$-wave symmetry. Fig. 3 shows the best fit to the tail of the resistive transition as a function of reduced temperature under rf irradiation of three different power levels. In calculation we adopted $`\mathrm{\Delta }(0)/k_BT_c^{^{}}=3.85`$ as obtained previously and assumed that $`\mathrm{\Delta }(T)`$ follows the BCS gap equation. We also set $`N(0)`$ to be 0.8 states/\[eV\]\[Cu-atom\] and adopt the nominal value for the effective volume of the superconducting electrode, $`\mathrm{\Omega }`$=25 $`\mu `$m$`\times `$35 $`\mu `$m$`\times `$0.3 nm. The inelastic scattering time $`\tau _i`$ is taken as the fitting parameter. The best fit values of $`\tau _i`$ depend on the rf irradiation power and turn out to be $`2.7\times 10^8`$ sec, $`2.0\times 10^8`$ sec, and $`4.6\times 10^9`$ sec for the power of -20 dBm, -23 dBm, and -30 dBm, respectively. $`\tau _i`$ gets longer as the transition temperature decreases for higher rf power. Fits below the superconducting transition with only one fitting parameter $`\tau _i`$ turn out to be reasonably good.
The values of $`\tau _i`$ determined in this way appear unreasonably long compared to the ones obtained from the $`R(T)`$ data above $`T_c`$ based on the two-dimensional superconducting fluctuation theory. For instance, the authors of Ref. estimated $`\tau _i`$ to behave as $`\tau _i=0.75\times 10^{13}`$(100 K/$`T)^2`$ sec for Bi2212 whiskers, which would give a few orders of magnitude shorter inelastic scattering time in the temperature range used in this experiment. As described below we believe this long inelastic scattering time stems from the two impeding mechanisms, both spatial and temporal, for effective charge-imbalance relaxation. In general, a charge-imbalanced nonequilibrium state relaxes to a charge-balanced nonequilibrium state, where pairs of conjugate electron- and hole-like quasiparticles relax to the paired ground state within the quasiparticle recombination time, $`\tau _{rec}`$, which is essentially given by $`\tau _r=3.7\tau _i[k_BT/\mathrm{\Delta }(T)]`$. In addition to this temporal relaxation, in conventional quasi-one-dimensional systems including a NS interface with homogeneous S electrode, any local nonequilibrium state in S can diffuse away within the characteristic length scale, $`\mathrm{\Lambda }_Q^{}`$, which is the spatial relaxation. In our highly anisotropic Bi2212 system, however, the spatial relaxation is hampered because the tunneling of qusiparticles in the surface CuO<sub>2</sub> layer to the neighboring inner CuO<sub>2</sub> layer is effectively blocked by the presence of the finite gap in the inner layer near $`T_c^{^{}}`$. In addition, the recombination rate of the quasiparticles within the surface CuO<sub>2</sub> layer becomes significantly slower as the gap shrinks as it approaches $`T_c^{^{}}`$. As a result of these two mechanisms the low-lying quasiparticle energy levels are highly occupied by the accumulating quasiparticles, blocking the further relaxation of the charge-imbalance nonequilibrium. Thus, in this situation, even the charge-balanced state is in nonequilibrium and the relaxation of the charge imbalance itself may slow down in the surface layer. It significantly enhances the charge-imbalance-induced RA in the surface layer and may explain the very long charge-imbalance relaxation time we obtained. In this sense, the quasiparticle recombination time is the bottleneck in the effective relaxation of the charge-imbalance nonequilibrium in the surface layer. Thus, one can conclude that what was actually measured in this experiment was the quasiparticle recombination time. In fact, our values from the fit above are in a reasonable range if we compare them with the quasiparticle recombination time, 0.8 ns and 2 $`\mu `$s, obtained in stacks of Bi2212 single crystals by Tanabe et al. and Yurgens et al., respectively, from fits to the back-bending of $`IV`$ characteristics, another nonequilibrium effect in dc measurements.
Although we take into account the $`d_{x^2y^2}`$-wave symmetry of the superconducting state of the surface CuO<sub>2</sub> layer by adopting angle-averaged quasiparticle density of states for the fit, the effect of the existence of nodes in the gap is not directly taken into account. No microscopic theory on the generation and the relaxation of the charge-imbalance nonequilibrium state for a NS interface where S has a $`d_{x^2y^2}`$ order-parameter symmetry is available to date. Naive consideration may lead to a conclusion that the exsistence of nodes makes the generation of the charge imbalance easier near the node region in the $`k`$ space. It may also slow down the relaxation process near the node region at any temperatures below the superconducting transition. Theoretical studies are required on the systems involving superconductors with $`d_{x^2y^2}`$-wave symmetry.
## IV ACKNOWLEDGMENTS
We appreciate the useful discussion with Prof. Soon-Gul Lee on the charge-imbalance phenomena. This work was supported by BSRI, MARC, KOSEF, and POSTECH. |
warning/0003/math-ph0003033.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Interest in wavelet theory and its applications in multi-resolution analysis has grown over the last decade, involving new and linking disparate fields of research from pure mathematics and physics to down-to-earth signal engineering. It is being widely applied in signal processing and data compression \[2-4\], pattern recognition , statistical physics and turbulence , jet dynamics , field theory , solid state physics \[10-12\], quantum mechanical applications \[13-16\], non-linear dynamics , soliton theory , etc.
The rise in interest in wavelet theory is motivated by the fact that Fourier analysis is ineffective when dealing with non-linear models and localize sharp features \[6-18\]. A central property of wavelets is their ability to expand and analyze functions with respect to a set of self-similar localized basis functions (scaling functions and wavelets) and to processes them locally without affecting the scale. The wavelet method is recursive and therefore ideal for computational applications. Moreover, the scaling functions and the corresponding wavelets are very well localized both in the time and frequency domains, and hence wavelets are ideal for an analysis of phenomena where different space/time scales occur. They also provide mathematical representations that can handle both analytical and numerical difficulties due to singular phenomena . Like Fourier analysis, wavelet theory uses basis functions with different characteristic scales. However, Fourier analysis has the advantage of being build upon a simple and solid mathematical foundation . The challenge is to create such a platform for wavelet theory.
This paper is a first step in that direction. It focuses on obtaining a closed algebraic method for the construction of the scaling and wavelet functions. Results for the Haar and B-wavelets are given. The results show that in addition to providing a multi-resolution basis of self-similar functions, wavelets display a definite non-linear symmetry which can be associated with a q-deformation of the Fourier series generating algebra. In a second paper we will show that wavelets also have variational properties: all multi-scale equations follow from Hamiltonโs equation of an infinite-dimensional Hamiltonian system.
The two-scale equation, which is the central object of wavelet theory, is also shown to have an algebraic structure that is associated with a quantum algebra . Applications of this result can be realized using supercomputers and efficient numerical schemes which, in turn, can be used to model complex continuous systems or genuine discrete systems defined on lattices. The symmetries related to complete solvability or to discrete-continuous transitions , and their basic tools, the finite difference equations, are new and key elements in any analysis of discrete systems. On uniform lattices the symmetry algebras of partial differential equations is left unchanged by the discretization . In the case of non-uniform lattices one needs to introduce generalized symmetries, like quantum algebras . This is an example of when the q-deformation of some initial symmetries can play an important role. An outcome of the present paper is an extension of traditional Fourier analysis towards wavelet expansions by means of q-deformation.
The finite-difference and scaling operators involved in the dilation equation are algebraically closed with respect to certain non-linear commutation relations. This symmetry is seen most simply when the operators are realized in terms of q-deformed derivatives . By expressing the translation and dilation operators as q-deformed derivatives, the scaling and difference operators can be mapped onto the generators of a non-linear algebra and the action of the q-deformed derivatives is extended to non-differentiable functions. Moreover, such non-linear algebras can be un-deformed to the Fourier series generating algebra or can be mapped onto the $`su_q(2)`$ algebra.
After introducing notation and basic definitions in Section 2, we present in Section 3 the deformation of a Fourier algebra into a scaling function generating algebra, and identify a spectral problem with the Haar dilation equation. In Section 4, we introduce the definition of the most general scaling function and wavelet algebra and prove that the algebraic formulation of the dilation equation is unique. A duality relation is obtained: for any scaling function there is a non-linear algebra; and for special algebras one can find a corresponding wavelet structure. In Section 5 we give examples of scaling function algebras for the Haar and the B-wavetes. We use new limiting procedures to solve the q-difference and finite-difference equations. In Section 6 we use a variant and extension of the deforming functional technique to obtain a mapping of the generators of non-linear algebras to those of $`su_q(2)`$. Concluding comments, further extensions, and other remarks are given in Section 7.
## 2 Basic elements
An algebraic structure can be built for any scaling function system and wavelet basis. In order to realize this construction, three basic building blocks are needed: the algebraic structure of the Fourier system (starting point), the wavelet theory (final objective), and q-deformation (intermediate tool).
### 2.1 Fourier algebras
In order to demonstrate our approach, consider algebras which have Fourier series as the basis of their representation spaces . In the following we shall denote $`^kf/x^k=_kf=f^k`$. The trigonometric (Fourier) system $`\{|k>=e^{ikx}\}_{kZ}`$ diagonalizes all translation invariant operators acting on $`L^2([0,2\pi ])`$. We introduce three generators within a differential realization $`J_0=i`$, $`J_\pm =e^{\pm ix}(i)^p`$, satisfying the commutation relations
$`[J_0,J_\pm ]=\pm J_\pm ,[J_+,J_{}]=1(12i)^p.`$ (1)
For $`p=0`$, eqs.(1) describe an algebra, denoted $`_0`$, that is isomorphic to an analytical prolongation, $`e(2,C)`$, of the Euclidean algebra $`e(2,R)`$ of rigid motions in the plane. The algebra $`e(2,R)`$, generated by $`P_x=P_x^{}`$, $`P_y=P_y^{}`$ and $`R=R^{}`$, with the commutators $`[R,P_{x,y}]=\pm P_{y,x}`$, $`[P_x,P_y]=0`$, is realized through the mapping
$$J_0=iR,J_\pm =P_x\pm iP_y,$$
onto the $`_0`$ algebra. The unitary irreducible representations (unirreps) of $`_0`$ are based on the eigenvectors of the self-adjoint operator $`J_0`$, $`J_0|n>=n|n>`$. For any two distinct eigenstates, $`|n>`$ and $`|n^{^{}}>`$, by using the first commutator in eq.(1), we obtain $`n^{^{}}=n+1`$ and hence the spectrum of $`J_0`$ is unbounded, discrete and consists in equidistant eigenvalues. Hence the space of representations is generated by the Fourier system, and the other generators $`J_\pm |n>=|n\pm 1>`$ act like ladder operators on the $`|n>`$ states, increasing/decreasing the scale by unity. For $`p=1`$ eqs.(1) describe another Lie algebra, $`_1`$, isomorphic with the symplectic algebra $`sp(2,R)su(1,1)so(2,1)`$. This differential realization has the same representation space as $`_0`$ but it is not irreducible. The action of the generators of $`_1`$ is similar to that of the generators of $`_0`$ for $`n0`$. Unlike the $`_0`$ case, however, we have that $`J_\pm |1>=|0>`$ and the sub-spaces $`\{e^{inx}\}_{nN}`$ and $`\{e^{inx}\}_{nN}`$ are invariant subspaces of $`_1`$. The $`_{0,1}`$ examples suggest the possibility of other constructions in terms of the operator $``$ and the complex exponential functions of different scales.
### 2.2 Scaling functions and wavelets
Wavelets are able to reconstruct a signal through regular sampling, namely, by analyzing the signal at different scales (which increase/decrease exponentially) with the step size between each scale being the same. Different from the ordinary Fourier transform, which reproduces a function as a superposition of complex exponentials, or from the windowed Fourier transform, which introduces a scale into the analysis of signals, multi-resolution analysis processes the signal locally, using the appropriate local scale. Basically, wavelets are constructed with a pair of operators: the dilation (scaling) and finite-difference (combinations of translations) operators, acting on $`L^2(R)`$ and defined, respectively, by $`T^\alpha f(x)=f(x+\alpha )`$, $`D^\beta f(x)=2^\beta f(2^\beta x)`$, with $`\alpha `$, $`\beta `$ arbitrary real numbers. They are invertible, unitary, and fulfill $`T^\alpha D^\beta =D^\beta T^{2^\beta \alpha }`$, $`T^\alpha T^\beta =T^{\alpha +\beta }`$, $`D^\alpha D^\beta =D^{\alpha +\beta }`$. The formal Taylor series representation of these operators are $`T^\alpha =e^\alpha `$ and $`D^\beta =2^\beta e^{\beta ln2x}`$. On the space of compact supported or rapidly decreasing functions, any holomorphic function $`f(T)`$ is locally polynomial. Indeed, if $`f`$ is holomorphic, and its action is taken on the compact supported function subspace of $`\mathrm{\Phi }L^2(R)`$, then the action of $`f(T)=_{kZ}C_kT^k`$ reduces to that of a Laurent polynomial by keeping only a finite number of terms in the sum.
The wavelet system is given by a set of scaled and translated copies of a pair of functions: the scaling function $`\mathrm{\Phi }`$ and the mother wavelet $`\mathrm{\Psi }`$. The basic fact about wavelets is that both these fundamental functions are finite linear combinations of $`\mathrm{\Phi }`$, reflecting the self-similar character of the wavelet system. The defining equation for the scaling function (the dilation equation) is a linear finite-difference equation, including a scale change (q-difference )
$`\mathrm{\Phi }(x)=D{\displaystyle \underset{k=0}{\overset{n}{}}}C_kT^k\mathrm{\Phi }(x)=Dg(T)\mathrm{\Phi }(x),`$ (2)
where the RHS sum is a polynomial in $`T`$, $`g(T)`$. Eq.(2) is a fixed-point equation and consequently it has only one unique solution \[1-7,11,13\]. The scaling function $`\mathrm{\Phi }(x)`$, as a solution of eq.(2), is required to have two properties \[1-5\]:
1. $`_R\mathrm{\Phi }(x)๐x=1`$, the average value property;
2. $`<\mathrm{\Phi }_n,\mathrm{\Phi }>_R\mathrm{\Phi }(x+n)\mathrm{\Phi }(x)dx=\delta _{n,0}`$ for any $`nZ`$, the orthogonality condition. These conditions introduce restrictions on the coefficients $`C_k`$ in eq.(2) \[1-4\],
$`{\displaystyle \underset{k=0}{\overset{n}{}}}C_k=2,{\displaystyle \underset{k=0}{\overset{n}{}}}C_kC_{k+2l}=\delta _{0l},lZ.`$ (3)
The conditions expressed through eq.(3) yield a pattern of $`L^2(R)`$ as a chain of subspaces $`V_jV_{j+1}`$, each one being generated by all the translations of $`D^j\mathrm{\Phi }`$, $`jZ`$. By repeated application of $`D`$ and $`T`$ on $`\mathrm{\Phi }`$, $`D^jT^n\mathrm{\Phi }(x)=\mathrm{\Phi }(2^jx+n)\mathrm{\Phi }_{j,n}`$, one obtains a non-orthogonal basis in $`L^2(R)`$. The action of the operator $`DT^\lambda g(T^1)`$ on $`\mathrm{\Phi }`$, $`\lambda `$ being a unique odd integer, provides the mother wavelet function:
$`\mathrm{\Psi }(x)=T^\lambda Dg(T^1)\mathrm{\Phi }(x)=T^\lambda {\displaystyle \underset{k}{}}(1)^kC_{k+1}\mathrm{\Phi }(2xk).`$ (4)
The wavelet $`\mathrm{\Psi }(x)`$ has the property that $`\{\mathrm{\Psi }_{j,n}\}_{j,nZ}`$ is an orthonormal basis of $`L^2(R)`$ \[1-3\] and $`L^2(R)`$ is a direct sum of the orthogonal subspaces $`W_j`$ (the orthogonal complements of $`V_j`$), each of them generated by all possible translations of $`D^j\mathrm{\Psi }=\mathrm{\Psi }_{j,0}`$ with integral m: $`L^2(R)=_{jZ}W_j`$, \[1-5\]. Hence translated and dilated copies of the mother wavelet, $`\mathrm{\Psi }(2^jx+k)`$, generate a true orthonormal basis. Recursively applications of the eqs.(2,4) relate all the scaling functions and wavelets.
The simplest example is provided by the Haar wavelet, defined by the scaling function $`\mathrm{\Phi }_{Haar}(x)=1`$ if $`|x1/2|1/2`$ and $`0`$ otherwise. In this case we have $`D(1+T^1)\mathrm{\Phi }_{Haar}=\mathrm{\Phi }_{Haar}`$ and $`g_{Haar}(T)=1+T^1`$ ($`C_0=C_1=1`$ and the rest $`0`$ in eq.(2)). The corresponding Haar wavelet is $`\mathrm{\Psi }_{Haar}=D(1T^1)\mathrm{\Phi }_{Haar}`$, i.e. $`\lambda =1`$ in eq.(4).
There are similarities and differences between the Fourier and the wavelet approaches. From an algebraic point of view, in both approaches, there are eigenfunction equations, which keep the scale constant, and the ladder operators, which change the scale. The dilation equation, eq.(2), has its analog in the Fourier formalism, though it is a fixed-point equation and has only one solution. The mother wavelet, eq.(4), has no analog in Fourier analysis. Each Fourier eigenfunction carries one scale; a wavelet function $`\mathrm{\Psi }(x)`$ involves two scales, e.g. $`\mathrm{\Phi }_{0,k}`$ and $`\mathrm{\Phi }_{2,k}`$. Because of their localization, wavelets possess a degree of freedom beyond that of a Fourier system; namely, the width of the support of $`\mathrm{\Phi }`$. In the present approach this degree of freedom is associated with the deformation parameter, $`q`$.
### 2.3 Quantum deformations
A possibility for constructing scaling functions/wavelets within an algebraic approach is to deform the Fourier algebraic structure into a non-linear system. In general, the scaling functions and wavelets are not differentiable. This suggests the use of finite-difference operators instead of derivatives. Finite-difference operators are closed with respect to commutation but only within non-linear algebraic constructions. Consequently, it is natural to use q-deformed derivatives to find a foundation for wavelet theory in $`q`$-deformed algebras.
Quantum algebras refer to some specific deformations of Lie algebras, to which they reduce when the deformation parameter $`q`$ is set equal to unity (for a recent monograph see and references therein). The simplest example of a $`q`$-algebra is su<sub>q</sub>(2) whose Jordan-Schwinger realization is given in terms of $`q`$-bosonic operators . The q-deformed algebras have been applied in various branches of physics like spin-chain models, non-commutative spaces, rotational spectra of deformed nuclei, Hamiltonian quantization, and dynamical symmetry breaking. The basic element is the q-deformation of a certain object $`x`$, which can be a number, an operator, or a function:
$$[x]_s=\frac{q^xq^x}{qq^1}\stackrel{s0}{}x,$$
(5)
where $`q=e^s`$. The Taylor expansion of $`T`$ and $`D`$ are related to the q-deformation of the derivative operator, according to the definition of the coordinate description of the q-deformed oscillator introduced in ,
$$[]_sf=\frac{T^sT^s}{2sinh(s)}f(x).$$
(6)
Analogously, we introduce the operator
$$[x]_{s\mathrm{ln}2}f(x)=\frac{f(2^sx)f(2^sx)}{2sinh(s\mathrm{ln}2)}=\frac{D^sD^s}{2sinh(s\mathrm{ln}2)}f(x).$$
(7)
When $`s0`$, $`[]_s`$ and $`[x]_{s\mathrm{ln}2}x`$. Eqs.(6,7) can be inverted and the occurrence of the q-deformed derivative and the translation/dilation operators becomes immediate
$$T^{\pm s}=\frac{1}{2}\left(\pm \frac{[]_s}{\eta (s)}+\frac{[]_{s/2}^2}{\eta ^2(s/2)}+2\right)$$
(8)
$$D^{\pm s}=\frac{1}{2}\left(\pm \frac{[x]_s}{\eta (s\mathrm{ln}2)}+\frac{[x]_{s/2}^2}{\eta ^2(s\mathrm{ln}2/2)}+2\right),$$
(9)
where $`sN`$ (or more general $`R`$) and $`\eta (s)=\frac{1}{e^se^s}`$. The q-deformed algebra $`su_q(2)`$, generated by $`\{J_0,J_\pm \}`$, is defined by the q-deformed version of the commutation relation
$$[J_+,J_{}]=[2J_0]_s$$
while the other two commutator relations remain undeformed, that is, they have the same form as for $`su(2)`$. In addition to this traditional version of $`su_q(2)`$, several generalized forms have been introduced through two different prescriptions: 1) by deforming the commutator $`[J_0,J_\pm ]`$ by using some arbitrary function $`G(J_0,q)`$, $`[J_0,J_+]=G(J_0,q)J_+`$ and $`[J_0,J_{}]=J_{}G(J_0,q)`$, independently proposed in , and ; and 2) by deformations involving all three commutation relations, using two functions $`G(J_0,q)`$ and $`F(J_0,q)=[J_+,J_{}]`$, introduced in . Unlike the former, for which the spectrum of $`J_0`$ is linear, the latter is characterized by an exponential spectrum for $`J_0`$. Since in wavelet theory the domains of analysis are divided exponentially, rather than linearly with bands of equal widths, such algebras are used in the following analyses.
## 3 Haar scaling function algebra, $`๐_{s,\alpha }`$
The aim in this section is to obtain operators depending on $`D`$ and $`T`$ that form a non-linear algebra. This structure must provide an algebraic form for eqs.(2,4) and map onto the Fourier algebra, $`_{0,1}`$. Consider an operator depending on $`T`$ and two real parameters $`s,\alpha `$:
$$W_0(s,\alpha )=\frac{T^{s\alpha }T^{s\alpha }\mathrm{cos}s\pi }{2\xi (\alpha )\mathrm{sinh}s},$$
(10)
where $`\xi (\alpha )=\frac{1}{\mathrm{sinh}(1)}\mathrm{sin}(\frac{\alpha \pi }{2})2i\mathrm{cos}(\frac{\alpha \pi }{2})`$. The $`(s,\alpha )`$ parameters allow $`W_0`$ to approach $``$ or combinations of $`T`$
$$W_0(0,\alpha )=\frac{\alpha }{\xi (\alpha )},$$
(11)
$$W_0(\frac{1}{2},\alpha )=\frac{1}{2\mathrm{sinh}(1/2)\xi (\alpha )}T^{\frac{\alpha }{2}},$$
(12)
$$W_0(1,\alpha )=\frac{T^\alpha +T^\alpha }{2\mathrm{sinh}(1)\xi (\alpha )},$$
(13)
$$W_0(2,\alpha )=\frac{T^{2\alpha }T^{2\alpha }}{2\mathrm{sinh}(2)\xi (\alpha )}.$$
(14)
In the limit $`s0,`$ $`W_0`$ reduces to the normal derivative with respect to $`x`$, eq.(11). In the limit $`s=\frac{1}{2}`$, we obtain a power of the translation operator $`T`$, eq.(12). Eq.(13) defines a q-deformation of unity, namely, $`W_0(1,0)=\frac{i}{4\mathrm{sinh}(1)}`$. $`W_0(s,\alpha )`$ can be a Hermitian or anti-Hermitian operator, $`W_0(1,\alpha )=\pm W_0(1,\alpha )^{}`$, depending on $`\alpha `$. The last limit, eq.(14), represents a finite-difference operator and is proportional to the q-derivative with respect to the $`\alpha `$-deformation $`[2]_\alpha `$, similar to eqs.(6-9). For $`\alpha =2k`$ in eq.(12), $`\alpha =k`$ in eq.(13) and $`\alpha =k/2`$ in eq.(14), the corresponding $`W_0(s,\alpha )`$ are Laurent polynomials in $`T`$, providing a direct connection to the wavelet operators. For some values of $`\alpha `$, eqs.(12-14) give invertible operators with respect to $`T`$, $`T=T(W_0)`$.
In order to introduce the dilation operator $`D`$, we define
$$W_\pm (s)=\frac{1}{2}e^{ix\frac{s1}{2}}D^se^{ix\frac{s1}{2}}(1+T^{\pm s}).$$
(15)
In order to fulfill the hermiticity condition $`(W_+)^{}=W_{}`$ on $`L^2(R)`$ these generators can be redefined in the form
$$W_\pm \stackrel{~}{W}_\pm =2^{1\frac{s}{2}}e^{ix\frac{s1}{2}}D^se^{ix\frac{s1}{2}}\left(1+e^{i\frac{s(s1)(1+2^s)}{4}}T^{\pm s(2^{\frac{s}{2}(11)})}\right).$$
By using the commutators relations between $`D`$ and $`T`$ and by performing an integration by parts, one has a direct check of the relation $`(\stackrel{~}{W}_+)^{}=\stackrel{~}{W}_{}`$
$$<f_1,\stackrel{~}{W}_+f_2>=_Rf_1^{}(x)\stackrel{~}{W}_+f_2(x)dx=<\stackrel{~}{W}_{}f_1,f_2>.$$
From eqs.(10,15) one obtains the commutators
$$[W_0(s,\alpha ),W_+(s)]=G(s,T)W_+(s),$$
(16)
$$[W_0(s,\alpha ),W_{}(s)]=W_{}(s)G(s,T),$$
(17)
$$[W_+(s),W_{}(s)]=F(s,T),$$
(18)
with
$$G(s,T)=W_0(s,\alpha )\frac{T^{2^ss\alpha }e^{is\alpha (1+2^s)\frac{s1}{2}}T^{2^ss\alpha }e^{is\alpha (1+2^s)\frac{s1}{2}}\mathrm{cos}s\pi }{2sinh(s)\xi (\alpha )}$$
(19)
and
$$F(s,T)=\frac{1}{4}\left((1+e^{is(1+2^s)\frac{s1}{2}}T^{2^ss})(1+T^s)(1+e^{is(1+2^s)\frac{s1}{2}}T^{2^ss})(1+T^s)\right).$$
(20)
Eqs.(10,15,16-20) describe a non-linear associative algebra, generated by $`T^\pm ,D^\pm `$, denoted $`๐_{s,\alpha }`$. The eigenvalue problem for $`W_\pm `$ provides the algebraic form for the scaling equation. When $`W_0(s,\alpha )`$ is invertible with respect to $`T`$, the algebra has $`W_0,W_\pm `$ as generators which depend on two parameters ($`s,\alpha `$) and is homomorphic with a special q-deformed algebra, namely, the two-color quasitriangular Hopf algebra $`๐^\pm `$ . The spectrum of $`W_0`$ is not equidistant because $`F(s,T)const.\times W_0`$. The Casimir operator of the algebra $`๐_{s,\alpha }`$ is given by \[28-31\]
$$C=W_{}W_++H(W_0)=W_+W_{}+H(W_0)F(W_0),$$
(21)
where $`H`$ is a real function, holomorphic in a neighborhood of $`0`$ and which must satisfy the functional equation
$$H(\xi )H(\xi G(\xi ))=F(\xi ).$$
(22)
with $`\xi `$ generic. The irreps of $`๐_{s,\alpha }`$ are labeled by the eigenvalues $`W_0|a>=a|a>`$. The commutator in eq.(16) can be written in the form
$$W_+W_0=(W_0G)W_+.$$
(23)
If $`|a>|a^{^{}}>`$ are eigenvectors of $`W_0`$, with corresponding eigenvalues $`a`$, $`a^{^{}}`$, we have from eq.(23) a non-linear recursion relation for all the eigenvalues
$$a^{^{}}=aG(a).$$
(24)
The algebra $`๐_{s,\alpha }`$ can be mapped into $`su(1,1)`$, the Lie algebra of a Fourier series, in the limit $`s0`$, $`\alpha 2`$: $`W_\pm e^{\pm ix}=J_\pm `$, $`W_0i=J_0`$, where we take these limits in the order $`๐_{s_0,\alpha _0}=lim_{\alpha \alpha _0}lim_{ss_0}๐_{s,\alpha }`$. The continuous mapping $`๐_{s,\alpha }๐_{0,2}su(1,1)`$ is an algebraic morphism and one can consider wavelets as q-deformed generalizations of the Fourier series, in the above sense.
As a first example, the Haar scaling function can be associated with a particular case of $`๐_{s,\alpha }`$, namely, $`๐_{1,1}`$. In the algebra $`๐_{1,1}`$ the eigenproblem for $`W_{}`$ provides the dilation equation for the Haar scaling function $`\mathrm{\Phi }(x)`$ defined by eq.(2) with $`g(T)=1+T^1`$. We have in this case $`G=W_02W_0^2+1`$, and $`W_\pm =\frac{1}{2}D^1(1+T^{\pm 1})`$ and the commutators
$$[W_0,W_+]=(W_02W_0^2+1)W_+,$$
$$[W_0,W_{}]=W_{}(W_0W_0^2+1),$$
(25)
$$[W_+,W_{}]=\frac{1}{4}\left(T^2(W_0)T^{1/2}(W_0)T^{1/2}(W_0)+T^1(W_0)\right),$$
where $`T(W_0)`$ is the solution of the equation $`2W_0=T+T^1`$, a pseudo-differential operator. The Casimir operator of $`๐_{1,1}`$ is a constant. Indeed, eq.(22) for the general Casimir operator is given by
$$H(W_0)H(2W_0^21)=F(W_0),$$
(26)
and has, in terms of the $`T=T(W_0)`$ operator, the form
$$\chi (T)\chi (T^2)=\frac{1}{4}\left(T^2T^{1/2}T^{1/2}+T^1\right),$$
(27)
where $`\chi (T)=H(\frac{T+T^1}{2})`$. The unique analytical solution for eq.(27) reads
$$\chi (T)=\frac{1}{4}(const.TT^{1/2}T^{1/2}),$$
(28)
which gives for the Casimir operator a constant.
The spectrum of $`W_0`$ consists of periodic functions satisfying the equation $`\mathrm{\Phi }(x+1)+\mathrm{\Phi }(x1)=2\mathrm{\Phi }(x)`$. This takes one back to Fourier analysis. The way to the Haar scaling function is to use the spectrum of $`W_\pm `$ since in $`๐_{1,1}`$ the eigenproblem for $`W_{}`$ gives the dilation equation for the Haar scaling function $`\mathrm{\Phi }(x)`$. Once $`\mathrm{\Phi }`$ is obtained from $`2W_{}\mathrm{\Phi }=\mathrm{\Phi }`$ the generator $`W_0`$ carries $`\mathrm{\Phi }`$ into an infinite sequence of functions $`W_0^n\mathrm{\Phi }`$. These functions are mutually orthogonal and generate the space $`V_0`$. The action of $`W_\pm `$ on $`\mathrm{\Phi }`$ yields $`W_0^{2^n}\mathrm{\Phi }`$. Consequently, $`V_0`$ is an invariant space for all the generators of the algebra $`๐_{1,1}`$.
The spectrum of $`W_0`$ depends on the values of the parameters $`s,\alpha `$. In the case of $`๐_{1,1}`$ this relation is $`a^{^{}}=2a^21`$. The corresponding representations are infinite-dimensional and the eigenvalues of $`W_0|_{1,1}`$ are $`a_n=cosh(2^n\stackrel{~}{a})`$, $`nZ`$ for any $`\stackrel{~}{a}`$. The corresponding eigenfunctions have the form $`|a_n>_\pm =e^{\pm 2^n\stackrel{~}{a}x}`$ which results in an exponential spectrum for $`a_n`$ similar to the sequence of scales in wavelet theory. This is a self-similar spectrum with respect to $`\stackrel{~}{a}`$, like the sequence of scales in wavelet theory. A part of this spectrum is shown in Fig. 1. The action of $`W_\pm `$ is:
$`W_\pm |a_n>_\pm ={\displaystyle \frac{1+e^{\pm 2^n\stackrel{~}{a}}}{2}}|a_{n\pm 1}>_\pm .`$ (29)
Another basis for a representation of $`๐_{1,1}`$ is given by $`\underset{ยฏ}{|k>}=e^{\frac{i\pi x}{k}}`$, $`kZ`$. We have the action
$$W_0\underset{ยฏ}{|k>}=\mathrm{cos}\frac{\pi }{k}\underset{ยฏ}{|k>},W_\pm \underset{ยฏ}{|k>}=\frac{1}{2}\left(1+e^{\pm \frac{i\pi }{k}}\right)\underset{ยฏ}{|2^{\pm 1}k>},$$
and on this basis $`(W_+)^{}=W_{}`$. There are two invariant spaces, $`\underset{ยฏ}{|2^k>}`$ and $`\underset{ยฏ}{|2^k>}`$, $`kN`$, with $`W_\pm \underset{ยฏ}{|0>}=1`$.
The same procedure can be followed for any set of the parameters $`s,\alpha `$. For example, if we choose the algebra $`๐_{2,1/2}`$ with $`W_0`$ defined in eq.(12), we obtain the spectrum of $`W_0`$ described by the recursion relation:
$$a^{^{}}=a\left(a+\sqrt{a^2+1}+\frac{1}{a+\sqrt{a^2+1}}\right),$$
(30)
which also gives a non-linear, unbounded, discrete representation of $`๐_{2,1/2}`$.
The last thing to prove is the uniqueness of the two-scale equation in the $`๐_{1,1}`$ scaling function generating algebra. The two-scale equation, eq.(2), in its algebraic form $`2W_{}\mathrm{\Phi }=\mathrm{\Phi }`$ is not unique in $`๐_{1,1}`$ if there exists an operator, similar with that occuring in the dilation equation, which commutes with $`W_{}`$. This operator should contain higher powers in $`T`$ and consequently is an element of $`U(๐_{1,1})`$.
Proposition 1: In $`U(๐_{1,1})`$ there exists a unique operator $`X=Dx(T^1)`$ such that $`[W_{},X]=0`$. The function $`x(\xi )`$ is integer, not polynomial, and unbounded, for generic $`\xi `$.
Proof: We take for $`X`$ a Laurent series $`x(T^1)=_{kZ}C_kT^k`$ then the condition $`[W_{},X]=0`$ results in a recursion relation for the coefficients $`C_k`$:
$`C_{2k+1}=C_kC_{2k1},C_{2k}=C_kC_{2k2},`$ (31)
for any $`kZ`$. Eqs.(31) have one trivial solution which reproduces $`W_{}`$, $`x(T^1)=C_1(1+T^1)`$ and only one other solution, with $`C_k=\pm C_1`$ or $`0`$, uniquely defined ($`C_2=C_{\pm 3}=C_4=C_5=C_6=C_7=C_8=C_{11}\mathrm{}=0`$, $`C_2=C_5=C_6=C_9=C_{10}=C_{10}=C_{11}=\mathrm{}=C_1`$ and $`C_1=C_4=C_7=C_8=C_9=C_{12}=C_{12}=\mathrm{}C_1`$, etc.). The sequence of non-zero coefficients is infinite, and $`\{C_k\}`$ is not a Cauchy sequence. (q.e.d.) We note that the $`๐_{1,1}`$ algebra is also a Hopf algebra, , defined by eqs.(25) and by the coproduct, counit and antipode in the form:
$$\mathrm{}T^\pm =T^\pm T^\pm ,\mathrm{}D^\pm =D^\pm D^\pm ,ฯต(T^\pm )=ฯต(D^\pm )=1,$$
$$S(T^\pm )=T^{},S(D^\pm )=D^{}.$$
All its generators are primitive elements.
## 4 General scaling and wavelet algebra
This algebraic approach for the Haar scaling function can be generalized to yield a non-linear algebra for any scaling function, and conversely, to find the dilation equation and scaling function for certain algebras. The procedure starts with an algebra generated by the dilation and translation operators and constructs, within this algebra, the two-scale equation. The generators of the $`๐_{s,\alpha }`$ algebra can be generalized as
$`W_0j_0(T),W_\pm j_\pm (D,T)=e^{ix\frac{s1}{2}}D^se^{ix\frac{s1}{2}}j(T^{\pm 1}),`$ (32)
with $`j_0(T)`$ and $`j(T)`$ being arbitrary functions of $`T`$ and $`s`$, holomorphic in a neighborghood of $`T=1`$, with their dependence on $`s`$ being such that in the limit $`s0`$, $`j_0(T)i`$, $`j(T)1`$. In this case the commutation relations, eqs.(16-20), become
$`[j_0,j_+]=\stackrel{~}{G}j_+,[j_0,j_{}]=j_{}\stackrel{~}{G},[j_+,j_{}]=\stackrel{~}{F}`$ (33)
with $`\stackrel{~}{G}`$, $`\stackrel{~}{F}`$ depending on $`T`$ through $`j_0,j`$, respectively,
$`\stackrel{~}{G}=\stackrel{~}{G}(s,T)=j_0(s,T)j_0(s,T^{2^s}e^{i(1+2^s)\frac{s1}{2}}),`$ (34)
$`\stackrel{~}{F}=\stackrel{~}{F}(s,T)=j(s,e^{i\frac{s1}{2}(1+2^s)}T^{2^s})j(s,T^1)j(s,T^{2^s}e^{i\frac{s1}{2}(1+2^s)})j(s,T).`$ (35)
Eqs.(33-35) define a non-linear algebra denoted $`๐_{j_0,j}`$ as a generalization of $`๐_{s,\alpha }`$. If the function $`j(T)`$ is invertible with respect to $`T`$, $`๐_{j_0,j}`$ can be expressed in terms of the generators $`j_0,j_\pm `$ only and then $`\stackrel{~}{G}(T)=๐ข(j_0)`$, $`\stackrel{~}{F}(T)=(j_0)`$. The algebraic morphism $`j_0(T)W_0`$ and $`j(T)1+T^s`$ provides the mapping $`๐_{j_0,j}๐_{s,\alpha }`$. Since the first two commutators of $`๐_{j_0,j}`$, eq.(33), do not depend on the function $`j(T)`$, and the third commutator in eq.(33) does not depend on the function $`j_0(T)`$, the closure condition for $`๐_{j_0,j}`$ is independent of the functions $`j_0(T)`$ and $`j(T)`$ and hence they can be choosen in a convenient way to provide any dilation equation in the form of the eigenproblem for $`j_{}`$. The algebraic closure conditions for $`๐_{j_0,j}`$, together with the definitions of $`j,j_0`$, require that
$`j_0(T^2)j_0(T)=๐ข(j_0(T^2)),`$ (36)
$`j(T^2)j(T^1)j(T^{1/2})j(T)=(j_0(T)).`$ (37)
Both eqs.(36,37) are non-linear, and in general, difficult to solve analytically. The unique solution for the dilation equation are $`j_\pm `$ since the Casimir operator of this algebra depends on $`T`$ only. Choosing $`g(T)=j(T)`$ yields the dilation equation in the form $`2j_{}\mathrm{\Phi }=\mathrm{\Phi }`$. In the limit of $`๐_{s,\alpha }`$ this provides again the Haar two-scale equation. The corresponding scaling function belongs to the basis of the representation of $`j_{}`$ with eigenvalue $`1/2`$. The remaining arbitrary function $`j_0`$ can now be selected to obtain the wavelet generating operator $`j_{}j_0\mathrm{\Phi }=\mathrm{\Psi }`$, in agreement with eq.(4). One can formally construct a basis of the representation for $`j_{}`$ with the functions $`|n>=(\mathrm{ln}T)^n\mathrm{\Phi }`$. The procedure is the following: for a given algebra $`๐_{j_0j}`$ (therefore given functions $`๐ข`$, $``$) solve eqs.(36,37) with respect to the functions $`j_0,j_\pm `$ and obtain the corresponding dilation equation in the form $`2j_{}\mathrm{\Phi }=\mathrm{\Phi }`$. A certain combination of generators produces the corresponding wavelet equation for $`\mathrm{\Psi }`$. Conversely, given a scaling/wavelet function, and consequently its dilation equation and $`g(T)`$, one can choose $`j_{}=Dg(T)`$ and then solves the equation $`Dg(T^1)=j_{}j_0`$ with respect to $`j_0`$. With $`j_0,j_\pm `$ known one can solve eqs.(36,37) with respect to $`๐ข,`$ to construct the algebra.
This algebraic approach is in some sense universal. What is modified is the specific realization of the non-linear algebra in terms of $`D`$ and $`T`$. Constraints are imposed by the two limiting approaches (two-scale equation and Fourier limit). Loosley speaking, the closure of the algebra provides the fixed-point dilation equation and the non-linearity of the algebra provides the exponential scaling.
We also note that $`V_0=\{_kC_kT^k\mathrm{\Phi }\}`$,that is, the space generated by all the integer translations of $`\mathrm{\Phi }`$, and any other $`V_j=D^jV_0`$, is not invariant to the action of $`j_{}`$. For example, the condition $`j_{}D\mathrm{\Phi }=_kC_kT^kD\mathrm{\Phi }`$ implies the existence of an operator containing $`D`$, whose commutator with $`j_{}`$ is a function of $`T`$ only. From the definition of the algebra this is impossible, and consequently this proves the above conjecture. The $`V_j`$ spaces form, by recursion, a basis in $`L^2(R)`$.
In order to prove the uniqueness of the two-scale equation in the general case, we again use Proposition 1, for $`W_{}j_{}=Dj(T^1)=D_{kZ}j_kT^k`$. We have to solve the commutation equation $`[Dj(T^1),X)]=0`$ for $`x(T)=_{kZ}X_kT^k`$, $`X=Dx(T)`$. This implies that the arbitrary functions $`x(\xi )`$ and $`j(\xi )`$ must fulfill the functional condition ($`\xi `$ a generic variable)
$$x(\xi ^2)j(\xi )=x(\xi )j(\xi ^2).$$
(38)
Eq.(38) does not carry any restriction with respect to the values of the functions between $`0`$ and $`1`$. Since $`Dj(T)`$ represents a scaling operator we have, according to the dilation equation, $`j(1)=1`$, $`j(1)=0`$ which implies $`x(1)=0`$. If $`\mathrm{\Phi }`$ is the corresponding scaling function for $`j_{}`$ then $`X\mathrm{\Phi }`$ is also an eigenfunction of $`j_{}`$, $`j_{}(X\mathrm{\Phi })=X\mathrm{\Phi }`$. In order for $`Dx(T)`$ to satisfy the average value property (1) in subsection 2.2, we impose the additional condition $`x(1)=1`$. The last condition require of for $`DX(T)`$, the orthogonality condition (2) of subsection 2.2, is equivalent with the condition \[1-4\]
$$|x(T)|^2+|x(T)|^2=1.$$
(39)
By introducing eq.(38) in eq.(39) we obtain $`X=j_{}`$ which provides the trivial indentity solution, and hence uniqueness.
## 5 Construction of the scaling and wavelet algebra
The link between the scaling function and wavelet, and the corresponding scaling algebra is supported by the solutions of eqs.(36,37). In the following we present a method for solving these equations, from the wavelets towards its algebra. The dilation equation eq.(2) has the form of a fixed-point equation. Therefore we must try to find its eigenvectors as limits of some functional sequences. The limits of these sequences should be compact supported or rapidly decreasing functions for two reasons: to obtain scaling functions with good localization, and to provide correct behaviour of the action of the operators $`f(T)`$. Since the scaling function will be expressed as a limit of a sequence, one has to look for operators which commute with the limit (are absorbed in the limit). We choose a test function $`\mathrm{}_0(x)`$ and a sequence of operators $`f_n`$ from the universal covering $`U(๐_{j_0,j})`$ of the scaling function algebra introduced in the preceding section, such that the limit $`\mathrm{\Phi }(x)=lim_n\mathrm{}f_n\mathrm{}_0(x)`$ exists in the weak topology and it provides the scaling function.
In the following we show how to construct the scaling algebra $`๐_{j_0,j}`$ from the dilation equation, written in the form $`j_{}(T,D,s)\mathrm{\Phi }=\mathrm{\Phi }`$. We have the following proposition in the framework of the algebra $`๐_{j_0,j}`$, for $`s=1`$:
Proposition 2 Let $`f_n(j_0)`$ be a functional sequence in $`U(๐_{j_0,j})`$, $`s=1`$, $`\mathrm{}_0(x)`$ a (test) function and $`\mathrm{\Phi }(x)`$ the scaling function. If the limit $`\mathrm{\Phi }(x)=lim_n\mathrm{}f_n(j_0)j_{}^n\mathrm{}_0(x)`$ exists, and $`lim_n\mathrm{}\frac{f_n(j_0๐ข(s,j_0))}{f_{n1}(j_0)}=1`$, then $`j_{}\mathrm{\Phi }(x)=\mathrm{\Phi }(x)`$.
Proof:
From the RHS of the first condition in the hypothesis and from eq.(33) we have
$$f_n(j_0)j_{}^n=j_{}f_n(j_0๐ข(s,j_0))j_{}^{n1}=j_{}\frac{f_n(j_0๐ข(s,j_0))}{f_{n1}(j_0)}f_{n1}(j_0)j_{}^{n1}.$$
It follows:
$$j_{}\mathrm{\Phi }=\underset{n\mathrm{}}{lim}j_{}f_n(j_0)\mathrm{}_0(x)=\mathrm{\Phi },$$
q.e.d. It follows from Proposition 2 that for a given dilation equation $`j_{}\mathrm{\Phi }=\mathrm{\Phi }`$, (both $`j_{}`$ and $`\mathrm{\Phi }`$ given) we can find out a sequence $`f_n(j_0)`$, the operator $`๐ข(s,j_0)`$, and a test function $`\mathrm{}_0(x)`$, such that these objects satisfy the hypothesis of Proposition 2. Then it follows that one can construct the scaling algebra, since with $`j_0`$ and $`๐ข(s,j_0)`$ found, $`(s,j_0)`$ results from the last commutator in eq.(33). In general, one starts with $`j_0`$ as an arbitrary function of $`T`$ which can be deformed into $`i`$ when mapping $`๐_{j_0,j}`$ to $`_{0,1}`$. As $`j_{}`$ is provided by the two-scale equation, $`j_{}\mathrm{\Phi }=\mathrm{\Phi }`$, other solutions are forbiden. If $`lim_n\mathrm{}f_n(j_0)=f_{\mathrm{}}(j_0)const.`$, this limit should commute with $`j_{}`$ which is forbidden by Proposition 1. This procedure closes the construction of the algebra $`๐_{j_0,j}`$.
In the following we give an algorithm for finding $`f_n(j_0)`$ and $`j_0(T)`$ and illustrate it with two examples. Since any mother scaling function is defined by the dilation equation ($`Dh(T)\mathrm{\Phi }=\mathrm{\Phi }`$ or $`j_{}\mathrm{\Phi }=\mathrm{\Phi }`$) it is natural to search for solutions by using a recursion algorithm. For instance, we can find a test function $`\mathrm{}_0(x)`$ and a triplet of operators $`A,B,C`$ such that the limit $`\mathrm{\Phi }(x)=Blim_n\mathrm{}A^n\mathrm{}_0(x)`$ exists, and in addition, $`CB=BA^k`$ for a finite positive integer $`k`$. Then, we have the property $`C\mathrm{\Phi }(x)=\mathrm{\Phi }(x)`$. Indeed, $`C\mathrm{\Phi }=CBlim_n\mathrm{}A^n\mathrm{}_0=Blim_n\mathrm{}A^{n+k}\mathrm{}_0=Blim_n\mathrm{}A^n\mathrm{}_0=\mathrm{\Phi }`$. From the dilation equation we know that $`C`$ should have the form $`C=Dc(T)`$ with $`c(T)`$ a function of $`T`$. One of the simplest choices is to use $`A=D`$ and $`B=b(T)`$. Then we have $`CB=Dc(T)b(T)=c(T^{1/2})b(T^{1/2})D=b(T)A`$; that is, $`k=1`$ and a restriction for the arbitrary functions $`c(T)`$ and $`b(T)`$ arises:
$`c(T^{1/2})b(T^{1/2})=b(T)`$ (40)
This equation is useful in both directions (algebraic $``$ dilation equation), since one can start with a given scaling function (given $`c(T)`$) and find the operator $`b(T)`$ involved in the algebra and conversely.
Now consider the algorithm for the Haar scaling function, $`c(T)=1+T^1`$. We look for solutions of eq.(40) as Laurent series for $`b(T)=_{kZ}b_kT^k`$. In this case eq.(40) has an unique solution, $`b(T)=const.\times (1T^1)`$. This gives again an unique solution for $`\mathrm{\Phi }=(1T^1)lim_n\mathrm{}D^n\mathrm{}_0`$. We stress that $`Db(T)`$ is exactly the operator which gives the Haar wavelet. Further, we can express $`\mathrm{\Phi }_{Haar}`$ in the form $`\mathrm{\Phi }(x)=H(x)H(x1)`$, where $`H(x)`$ is the Heaviside distribution. For any sequence of $`C^{\mathrm{}}`$ functions $`\delta _n(x)\delta (x)`$, we have $`\mathrm{\Delta }_n(x)=\delta _n(x)๐xH(x)`$ and we can write
$$\mathrm{\Phi }(x)=\underset{n\mathrm{}}{lim}(\mathrm{\Delta }_n(x)\mathrm{\Delta }_n(x1)).$$
(41)
For example, we can use the sequences $`\delta _n=\frac{1}{\pi }\frac{2^n}{x^2+2^{2n}}`$ or $`\delta _n=\frac{2^{n1}}{cosh^2(2^nx)}`$ and $`\mathrm{\Delta }_n=\frac{arctan(2^nx)}{\pi }`$ or $`\mathrm{\Delta }_n=\frac{1}{2}tanh(2^nx)`$, respectively. The latter example is a soliton-like shape, having good localization. These sequences converge to $`\delta (x)`$, respectively. In this way we have selected subsequences that step in powers of $`2`$. Eq.(41) can be written as:
$`\mathrm{\Phi }(x)=\underset{n\mathrm{}}{lim}(1T^1)D^n\mathrm{}_0(x),`$ (42)
We can express this definition in terms of the algebra $`๐_{1,1}`$. By using eq.(25) and the properties of the operators $`D`$ and $`T`$, we can write eq.(42) in the form
$$\mathrm{\Phi }(x)=\underset{n\mathrm{}}{lim}(1T^{2^n})(2W_{})^n\mathrm{\Delta }_0(x).$$
(43)
Indeed, from the commutation relation between $`D`$ and $`T`$ and the definition of $`W_{}`$ we have
$`(2W_{})^n=(D(1+T^1))^n=D^n(1+T^1+\mathrm{}+T^{2^n+1})`$ (44)
$$=(1+T^{2^n}+(T^{2^n})^2+\mathrm{}+(T^{2^n})^{2^n1})D^n.$$
Hence, we can write $`(1T^1)D^n=(1T^{2^n})(2W_{})^n`$. Moreover, we can write, by using the inverted form for $`W_{}(T)`$, the dilation equation in a pure algebraic form:
$$\mathrm{\Phi }(x)=\underset{n\mathrm{}}{lim}(1T^{2^n}(W_0))(2W_{})^n\mathrm{\Delta }_0(x).$$
(45)
From this last equation it follows that $`2W_{}\mathrm{\Phi }=\mathrm{\Phi }`$, i.e., the Haar dilation equation. If the function $`\mathrm{\Delta }_0(x)`$ is choosen from a class of suitable functions for a wavelet analysis then its scaling function $`\mathrm{\Phi }(x)`$, eq.(45), gives a rapidly convergent wavelet expansion. In order to demonstrate this procedure, we show in Fig. 2 scaling functions obtained in this way for different values of the parameter $`s`$ within $`๐_{s,\alpha }`$. For $`s=0`$ and $`1`$ this functions yields the Haar scaling function.
We give another example for the $`\mathrm{\Phi }_2`$ B-scaling function and the $`\mathrm{\Psi }_2`$ B-wavelet. We define
$`\mathrm{\Phi }_2(x)=4x\text{for}0<x<1/2;4x+4\text{for}1/2<x<1,`$ (46)
and 0 in the rest, or
$`\mathrm{\Phi }_2(x)=\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{2^{n+1}}}(12T^{1/2}+T^1)D^n\mathrm{}_{0,tri}(x),`$ (47)
with
$`\mathrm{}_{0,tri}(x)={\displaystyle \frac{2x}{\pi }}arctan(2x){\displaystyle \frac{\mathrm{ln}(1+4x^2)}{2\pi }}={\displaystyle _0^{2\pi }}\mathrm{}_0(x)๐x`$ (48)
Following the above algorithm we obtain the corresponding dilation equation
$`j_{}|_{tri}\mathrm{\Phi }_2(x)={\displaystyle \frac{D}{2}}(1+T^{1/2})^2\mathrm{\Phi }_2(x).`$ (49)
By using the commutation relation $`D=2D`$ and the dilation equation eq.(49) we can obtain a self-contained equation for the wavelet:
$`\mathrm{\Psi }_{tri}(x)=D(1+2T^{1/2}+T^2)^2\mathrm{\Psi }_{tri}.`$ (50)
The corresponding algebras for Haar and $`\mathrm{\Phi }_2`$ B-scaling functions are different. For the B-wavelet we have, in $`๐_{1,1}`$
$`G(W_0)=W_0W_0^2{\displaystyle \frac{1}{2}}`$ (51)
This new algebra fulfills different commutation relations and consequently has a spectrum for $`W_0`$ which is different from that for the Haar scaling algebra. The algorithm introduced above can also be used to generate other scaling functions and their corresponding algebras. An interesting track is to obtain polygon-like wavelets or, by following the same procedure which guided us to obtain the B-scaling function from the Haar scaling function, to obtain smoother scaling functions.
## 6 The $`su_q(2)`$ limit of $`๐_{j_0,j}`$
The quantum group $`su_q(2)`$, or its extensions \[21-26,28-31\], are associative algebras over $`C`$ generated by three operators, $`J_0=(J_0)^{}`$, $`J_+`$, and $`J_{}=(J_+)^{}`$, satisfying the commutation relations
$$[J_0,J_+]=J_+,[J_0,J_{}]=J_{},[J_+,J_{}]=[J_0]_s,$$
(52)
where $`[J_0]`$ is a deformation of $`J_0`$, that is, a real parameter-dependent function of $`J_0`$, holomorphic in the neighbourhood of zero, and approaching $`2J_0`$ in the limit $`s0`$. We have to find a deforming functional that transforms the $`๐_{j_0,j}`$ generators into operators satisfying the commutation relations eq.(52). For this the first equation in eqs.(33) can be written as
$$\left(j_0๐ข(s,J_0)\right)j_+=j_+j_0,$$
(53)
and hence, for every entire function $`p(\xi )`$, we can write
$$p\left(j_0๐ข(s,j_0)\right)j_+=j_+p(j_0).$$
(54)
Let us consider the functional equation
$$\mathrm{\Gamma }\left(\xi ๐ข(s,\xi )\right)=\mathrm{\Gamma }(\xi )1$$
(55)
for a given function $`\mathrm{\Gamma }(\xi )`$ with $`\xi `$ generic. If this equation has a solution $`\mathrm{\Gamma }(\xi )`$ that is an entire function, then eq.(55) can be written, for $`p(\xi )=\mathrm{\Gamma }(\xi )`$, in the form $`[\mathrm{\Gamma }(j_0),j_+]=j_+`$, and correspondingly $`[\mathrm{\Gamma }(j_0),j_{}]=j_{}`$. These equations give the correspondence between the wavelet algebras and the q-deformed algebras like $`su_q(2)`$ or other extensions of them. They map the structure of $`๐_{j_0,j_\pm }`$, eq.(33), onto the structure described by the first two equations in eq.(52). For the third commutator of each of these algebras we use the function $`\mathrm{\Gamma }(\xi )`$, which allows eqs.(33) to be reduced to the third equation in eqs.(52) through the mapping:
$$J_0=\mathrm{\Gamma }(j_0),J_\pm =j_\pm ,$$
(56)
If $`\mathrm{\Gamma }`$ is invertible, the third equation in eqs.(33) in $`๐_{j_0,j}`$ can be reduced to the third equation of eqs.(52) of an algebra defined by the function $`[J_0]`$ with the identification $`\mathrm{\Gamma }^1=[]`$ where $`[]\mathrm{\Gamma }`$ means the composition of the two functions, i.e., $`([]\mathrm{\Gamma })(\xi )=[\mathrm{\Gamma }(\xi )]`$. In the case of su<sub>q</sub>(2), i.e. $`[J_0[=[2J_0]_q`$, the function $`(s,j_0,s)`$ becomes:
$$(s,j_0)=\frac{\left(\varphi (j_0)\right)^2\left(\varphi (j_0)\right)^2}{qq^1},$$
(57)
where $`q=e^s`$, $`\varphi (\xi )q^{\mathrm{\Gamma }(\xi )}`$ has to satisfy the equation $`\varphi \left(\xi ๐ข(s,\xi )\right)=q\varphi (\xi )`$. Consequently, eqs.(53-55) provide the reduction of $`su_q(2)`$ into $`๐_{j_0,j}`$ through $`\mathrm{\Gamma }`$. For details of the above technique of mapping and reduction of non-linear algebras one can see examples in . In the case of the algebra $`๐_{1,1}`$, eqs.(53-55) give the condition
$$p(W_01)=W_0G(1,W_0),$$
(58)
which has the solution
$`J_0=p(W_0)|_{๐_{1,1}}=cosh(2^{W_0}b),`$ (59)
for any arbitrary constant $`b`$. By inverting eq.(59) we obtain
$$W_0=\frac{1}{ln(2)}ln(\frac{}{b}).$$
(60)
Conversely, starting from $`su_q(2)`$ and its deformation $`[]_q`$ one can map it into $`๐_{1,1}`$ where for the operator $`F(s,T)=F(s,T(W_0))`$ we have
$$F=f(\mathrm{\Gamma }(W_0))=f(p^1(W_0))=[2p^1(W_0)].$$
(61)
A diagram containing these algebraic morphisms is presented in Fig.3.
## 7 Further extensions, comments and conclusions
In this paper we present a method for connecting classes of non-linear (q-deformed) algebras with scaling functions and wavelets. The connection is reciprocal. The algebra $`๐_{j_0,j}|_{s=1}`$ allows the choice of different dilation equations, i.e., for different functions $`j(1,T)`$ one can obtain different scaling functions. The function $`j_0(1,T)`$ should be choosen in order to give the correct wavelet function through the action of $`j_0j_{}`$. The connection with quantum groups is more transparent if we choose for $`j_0=[]_{s=1/2}=W_0(2,1/2)`$, eq.(14), and for $`\mathrm{\Phi }`$, the Haar scaling function, $`j_{}=D(1+T^1)`$. The action of the q-derivative $`[]_{1/2}`$ on $`\mathrm{\Phi }`$ gives exactly the Haar wavelet. In this example the operator $`j_0`$ is anti-Hermitian and we can identify its spectrum (imaginary eigenvalues) from the commutation relations. We also have the relation $`\mathrm{\Psi }=(j_0\stackrel{~}{G}(1,j_0)\mathrm{\Phi }`$, so that one can express the eigenvectors $`|a>`$ of $`J_0`$ in the wavelet basis of the algebra. If $`|a>=_{j,n}C_{j,n}\mathrm{\Phi }_{j,n}`$, we obtain a recursion relation for the coefficients of $`|a>`$ and for the eigenfunctions of $`j_0`$ $`_kC_{j,p2^jk}๐_k=aC_{j,p}`$ for any $`j`$ and $`p`$ integer, where $`๐_k`$ are the Taylor coefficients of the function $`\xi ^\lambda g(\xi )`$. We note that by using the q-derivative instead of a function of T we obtain exactly the same mother wavelet (Haar wavelet).
We note a natural extension of the above developments coming from a realization of $`sp(2,R)`$ in terms of the full algebra of the symmetry of the real line: $`X_0=x`$ (dilations), $`X_{}=`$ (translations), and $`X_+=x^2`$ (expansions). This is the maximal finitely generated real Lie algebra with generators in the form $`x^n`$. By exponentiation the generators $`X_0,X_{}`$ provide $`D`$ and $`T`$ and the last generator has the action $`E^a=e^{aX_+}f(x)=f(x/1ax)`$. For $`a>0`$ the action corresponds to a contraction of the function and for $`1<a<0`$ it is a dilation. But for $`a<1`$ and $`|a|1`$ we obtain a splitting of the function into two Heaviside distributions: $`E^af(x)=H(x)+H(x+1/a)`$. A possible closed non-linear algebra containing $`D,T`$ and $`E`$ might be an opportunity for the application of wavelets with multi-scale properties .
In conclusion, we have found that the operators of dilation and translation ($`D`$,$`T`$) can be combined in such a way as to generate non-linear algebras which depend on certain parameters ($`s`$,$`\alpha `$). We have investigated these algebras from the point of view of quantum groups, discussing their unirreps, Casimir operators, and reductions of these algebras to other q-deformed algebras, like $`su_q(2)`$ or to the Fourier series generating algebra. It has been shown that such algebras provide an appropriate framework for the foundation of wavelet analyses and for the obtaining the corresponding scaling functions. We have worked out two examples: the Haar and the B-scaling functions. The algorithm provides a general algebraic method for finding a specific scaling function. Direct applications of such an approach can be found in the theory of finite-difference equations and q-difference equations \[21-25,27\].
This study represents only a first step towards understanding the relation between the theory of non-linear algebras (having exponential spectra) and wavelets with their finite-difference equations. We conjecture that such an approach to scale invariant structures can lead to interesting mathematical constructions and tools for identifying isolated coherent structures associated with certain mother wavelets and limiting non-linear algebraic structures. The transition between such self-organized structures can be carried out through the modification of the deformation parameter $`q`$, with the intermediate domain representing โnoiseโ (non-closed algebraic structures).
FIGURE CAPTIONS
1. The dependence on $`a_0`$ of three eigenvalues $`a_n`$ of $`W_0`$ in $`๐_{1,1}`$, for $`n`$ = 2 (the largest period), 4 and 6 (the smallest period). The three spectra have self-similar structure for $`a_0[0,1]`$.
2. The continuous deformation of the scaling function $`\mathrm{\Phi }`$ in $`๐_{s,\alpha }`$ as a function of the parameter $`s`$. For $`s=0,1`$ this is the Haar scaling function.
3. A diagrammatic representation of the algebraic maps (arrows) and morphisms ($``$) of $`๐_{j_0j}`$, $`๐_{s,\alpha }`$, $`su_q(2)`$, $`su(2)`$ and $`_{0,1}`$. |
warning/0003/hep-th0003050.html | ar5iv | text | # 1. Introduction
## 1. Introduction
The AdS/CFT correspondence has improved considerably our understanding of strongly coupled large $`N`$ gauge theories , showing that the natural description of 4d theories in this regime is in terms of 5d gravity, where the extra dimension is related to the 4d energy scale. So far, strong evidence for the validity of the conjecture has come from the description of $`๐ฉ=4`$SYM in terms of IIB supergravity, while limited efforts have been devoted to the holographic description of non-supersymmetric non-conformal field theories. Moreover, the few examples analyzed were mostly based on the tachyonic (oriented) 0B string , where the presence of the tachyon complicates the gravity description and leads to instabilities in the dual strongly coupled field theory .
The first attempt to find a non-tachyonic holographic description of non-supersymmetric gauge theories was recently made in . The analysis was still performed within the framework of 0B strings, but resorting to a particular orientifold projection to eliminate the tachyon from the bulk and to avoid problems with the doubling of R-R sectors. Actually, the gauge theories that naturally emerge from (orientifolds of) type 0 models have the peculiar property of becoming conformal in the planar limit, and therefore several results can be mutuated from $`๐ฉ=4`$. In particular, the leading (planar) geometry is known to be of the form $`AdS_5\times X_5`$ with a constant dilaton field. For finite $`N`$, however, the gauge theory is no longer conformal and new features are expected in the gravity description. Indeed, a general property of non-supersymmetric theories seems to be the presence of a non-vanishing dilaton tadpole at genus one-half . On the other hand, supersymmetry relates R-R tadpoles to NS-NS ones, and thus demanding neutral configurations of orientifold planes and D-branes automatically ensures the vanishing of all massless tadpoles. However, whenever supersymmetry is not present one can find configurations for which the R-R and NS-NS tadpole conditions are incompatible . While uncanceled R-R tadpoles manifest themselves in pathologies of the string theory , NS-NS tadpoles are less problematic. They correspond to potential terms in the low-energy action and their main effect is to modify the vacuum. In our case this is precisely what one needs! Indeed, studying D3 branes in non-tachyonic orientifolds of type 0B (with $`\mathrm{O}^{}7`$-planes and D7-branes, where a prime indicates the non-tachyonic involution) we showed explicitly how a genus one-half dilaton tadpole in the bulk theory does reproduce qualitatively and quantitatively the expected logarithmic gauge theory RG flow. (Recently, a logarithmic running of the gauge coupling was also found in the gravitational description of $`๐ฉ=2`$$`\mathrm{SU}(N)\times \mathrm{SU}(N+M)`$ gauge theory . See related discussion on dilaton-induced RG flow in .)
The purpose of this letter is to pursue this program further, studying the back-reaction on the metric induced by the presence of a dilaton tadpole. In particular, we compute the Wilson loop in string theory and find a quark anti-quark potential interpolating between a (logarithmically running) Coulomb phase and a confining phase in the IR, as expected from the field theory analysis. To the best of our knowledge, the interpolating solution that we find is the only known example in which a single solution describes the expected behavior in the entire energy regime. Although our analysis is quite general and applies to a vast class of non-supersymmetric non-conformal theories for which a non-vanishing dilaton tadpole is expected to develop , the simplest model one may conceive originates from D3-branes and $`\mathrm{O}^{}3`$-planes in type 0B. In the large $`N`$ limit the near horizon geometry is $`AdS_5\times \text{โโ}^5`$ , where $`\text{โโ}^5=S^5/\text{}_2`$, with $`\text{}_2`$ the non-tachyonic involution associated to $`\mathrm{O}^{}3`$. Since the $`\text{}_2`$ acts freely on $`S^5`$, no open sector is present in the bulk and the spectrum of the model consists of all the $`\text{}_2`$-invariant harmonics of the 0B fields. In particular, the tachyon is projected out. Moreover, since two-cycles of $`\text{โโ}^5`$ are unorientable, as for instance $`\text{โโ}^2\text{โโ}^5`$, new interactions in the low-energy action are expected to originate from unoriented world-sheet topologies. These are weighted by generic (not only even) powers of the dilaton, and in particular a dilaton tadpole can be generated. Unfortunately, an explicit computation of one-point amplitudes for the dilaton in the curved $`AdS_5\times \text{โโ}^5`$ geometry seems out of reach within the actual (perturbative) definition of string theory, although such a term is expected from the above arguments, from our experience with non-supersymmetric theories, and from the results we are going to present here.
The organization of this letter is as follows: in section 2 we introduce the class of field theories we want to study, and present an explicit example from (unoriented) D-branes in the 0B string. In section 3 we turn to the gravity description and find an explicit solution for the dilaton and the metric up to next-to-leading order in $`1/N`$ expansion. In section 4 we extract the holographic quark anti-quark potential and compare it with the field theory analysis of section 2. Finally, section 5 contains our final comments about the validity of the gravity description.
## 2. Field theory considerations
One of the most remarkable results in field theory is the running of coupling constants
$$\frac{d\alpha (u)}{d\mathrm{log}u}=\beta (\alpha ),$$
(1)
where $`\alpha (u)=g_{\mathrm{YM}}^2(u)N`$ is the โt Hooft coupling. The precise expression of the beta-function is model dependent and encodes the quantum corrections to the classical behavior $`\alpha (u)=\alpha _0`$. Whenever a quantum theory does not depend on any scale and thus is exactly conformally invariant, as for instance $`๐ฉ=4`$SYM, $`\beta (\alpha )`$ vanishes identically and the coupling constants take their classical values for all energy scales. Combining these two arguments, it is then evident that theories that are conformally invariant in the planar limit are governed by a RG equation
$$\frac{d\alpha (u)}{d\mathrm{log}u}=\frac{1}{N}\beta _1(\alpha )\frac{1}{N^2}\beta _2(\alpha )+\mathrm{},$$
(2)
so that one recovers the expected $`\alpha (u)=\alpha _0`$ result for $`N\mathrm{}`$. Here, $`\beta _i(\alpha )`$ do not represent the $`i^{\mathrm{th}}`$-loop contribution to the RG equation. Rather, they encode all the contributions at a given order in $`1/N`$.
To be concrete, the simplest prototype one can conceive is an $`\mathrm{SU}(N)`$ gauge theory with six real scalars in the adjoint and 4 Weyl spinors in the antisymmetric and antisymmetric-conjugate representations, arising from the non-tachyonic orientifold projection on type 0B D3 branes . This theory was shown to converge to the $`\mathrm{SU}(N)`$ $`๐ฉ=4`$SYM in the planar limit , and to have a one-loop beta-function suppressed by $`1/N`$ (with $`b_1=\frac{16}{3}`$), and a two-loop beta-function contributing both to $`\beta _1(\alpha )`$ and $`\beta _2(\alpha )`$ in (2) .
One can extract some interesting results from the RG equation (2). Since in (2) there is a natural small parameter, $`1/N`$, it is natural to solve this equation by iterations
$$\alpha (u)=\alpha _0+\frac{1}{N}\alpha _1+\mathrm{}.$$
(3)
The solution
$$\alpha (u)=\alpha _0\frac{1}{N}\beta _1(\alpha _0)\mathrm{log}u$$
(4)
that, to this order, can be equivalently written as
$$\frac{1}{\alpha (u)}=\frac{1}{\alpha _0}+\frac{1}{N}\frac{\beta _1(\alpha _0)}{\alpha _0^2}\mathrm{log}u,$$
(5)
leads then to a quark anti-quark potential
$$E(L)=\frac{\alpha (L)}{L}=\frac{\alpha _0}{L(1\frac{1}{N}\frac{\beta _1(\alpha _0)}{\alpha _0}\mathrm{log}L)},$$
(6)
as expected for asymptotically free theories. Actually, since for large $`N`$ the beta-function varies slowly, the logarithmic behavior (5) persists beyond the perturbative regime, up to energy scales such that $`\left|\frac{1}{N}\frac{\beta _1(\alpha _0)}{\alpha _0^2}\mathrm{log}u\right|1`$. Note that the expansion is not valid in the (extreme) IR.
The description of the (non-perturbative) IR regime is less straightforward. Although for the specific model we introduced previously the scalars do not acquire a potential at the one-loop level , in general one would not expect to find a moduli space of vacua for non-supersymmetric theories. Thus, a potential for the scalars should develop: the scalars might become massive, or tachyonic, and might also acquire a non-vanishing vacuum expectation value. Although the specific shape of the potential is model dependent, in the following we shall assume that the scalars become massive without acquiring any vev. As we shall see, the string theory description supports this assumption.
Once the massive scalars decouple, the IR degrees of freedom will consist of Yang-Mills theory with four massless spinors (a QCD like theory), and a confining potential is expected
$$E(L)=\sigma L,$$
(7)
where $`\sigma =\mathrm{\Lambda }_{\mathrm{QCD}}^2`$ is the QCD string tension. For pure (large $`N`$) Yang-Mills theory one expects a constant (finite) string tension $`\sigma _0`$. In the present case, however, one expects a different behavior. Since the theories we are describing are conformal in the planar limit, the string tension should to vanish accordingly. To estimate the $`N`$ dependence of the string tension, one can integrate the RG equation (2) from an arbitrary weak coupling scale ($`u_0`$ in the UV) to a strong coupling one ($`\mathrm{\Lambda }_{\mathrm{QCD}}`$):
$$\mathrm{\Lambda }_{\mathrm{QCD}}=u_0e^{N{\scriptscriptstyle {\scriptscriptstyle \frac{d\alpha }{\beta _1(\alpha )}}}},$$
(8)
that translates into an exponentially suppressed string tension
$$\sigma e^N.$$
(9)
## 3. The gravity solution
We now turn to a dual gravitational description of the gauge theories introduced in the previous section, in the spirit of the AdS/CFT correspondence . As discussed in , the dual gravity description involves the ten-dimensional metric, the dilaton and the R-R four-form potential (with self-dual field strength), whose dynamics is essentially encoded in the action<sup>3</sup><sup>3</sup>3To be more precise one should use the PST action , since we are dealing with a self-dual 5-form $`F_5`$.
$$S=\frac{1}{(\alpha ^{})^4}d^{10}x\sqrt{g}\left[e^{2\mathrm{\Phi }}(R+4_\rho \mathrm{\Phi }^\rho \mathrm{\Phi })+\frac{1}{\alpha ^{}}Ce^\mathrm{\Phi }\frac{1}{4}(\alpha ^{})^4|F_5|^2\right],$$
(10)
or in its equations of motion
$`e^{2\mathrm{\Phi }}\left(8(\mathrm{\Phi })^28^2\mathrm{\Phi }2R\right)Ce^\mathrm{\Phi }`$ $`=`$ $`0,`$ (11)
$`e^{2\mathrm{\Phi }}\left(R_{\mu \nu }+2_\mu _\nu \mathrm{\Phi }\right)+\frac{1}{4}g_{\mu \nu }Ce^\mathrm{\Phi }`$
$`\frac{1}{96}(F_{\mu \rho \sigma \lambda \tau }F_\nu {}_{}{}^{\rho \sigma \lambda \tau }\frac{1}{10}g_{\mu \nu }F_{\rho \sigma \lambda \tau \eta }F^{\rho \sigma \lambda \tau \eta })`$ $`=`$ $`0.`$ (12)
The term $`Ce^\mathrm{\Phi }=Cg_s^1`$ corresponds to the dilaton tadpole expected from half-genus in string perturbation theory. Although the direct computation of this term is difficult, there are plausible arguments in favor of its presence, as discussed in the introduction.
In the spirit of AdS/CFT correspondence, we make for the metric the ansatz
$$ds^2=\alpha ^{}\left(d\tau ^2+e^{2\lambda (\tau )}\eta ^{ab}dx_adx_b+e^{2\nu (\tau )}d\mathrm{\Omega }_5^2\right),$$
(13)
compatible with four-dimensional Poincarรฉ invariance and SU(4) flavor symmetry. Moreover, the five-form field strength has $`N\frac{1}{2}`$ units of flux induced by the $`N`$ D3 branes and the O$`{}_{}{}^{}3`$ plane. In the following we shall ignore the contribution of the orientifold plane, manifestly suppressed in the presence of a large number of branes.
Inserting the above ansatz in the eqs. (11) and (12) and allowing for a $`\tau `$ dependence of the various fields one obtains
$`2\ddot{\mathrm{\Phi }}4\ddot{\lambda }5\ddot{\nu }4\dot{\lambda }^25\dot{\nu }^2+\frac{1}{2}N^2e^{2\mathrm{\Phi }10\nu }\frac{1}{4}Ce^\mathrm{\Phi }`$ $`=`$ $`0,`$ (14)
$`\ddot{\lambda }(2\dot{\mathrm{\Phi }}4\dot{\lambda }5\dot{\nu })\dot{\lambda }\frac{1}{2}N^2e^{2\mathrm{\Phi }10\nu }+\frac{1}{4}Ce^\mathrm{\Phi }`$ $`=`$ $`0,`$ (15)
$`\ddot{\nu }(2\dot{\mathrm{\Phi }}4\dot{\lambda }5\dot{\nu })\dot{\nu }4e^{2\nu }+\frac{1}{2}N^2e^{2\mathrm{\Phi }10\nu }+\frac{1}{4}Ce^\mathrm{\Phi }`$ $`=`$ $`0,`$ (16)
$`12\dot{\lambda }^220\dot{\nu }^24\dot{\mathrm{\Phi }}^2+20\dot{\mathrm{\Phi }}\dot{\nu }+16\dot{\mathrm{\Phi }}\dot{\lambda }40\dot{\lambda }\dot{\nu }`$
$`+20e^{2\nu }N^2e^{2\mathrm{\Phi }10\nu }+Ce^\mathrm{\Phi }`$ $`=`$ $`0.`$ (17)
As usual in General Relativity, the $`\tau \tau `$ component (17) of Einsteinโs equations is not independent, and translates into a vanishing energy constraint for the associated mechanical system.
In order to better appreciate the role of the different contributions in the low-energy action, it is useful to redefine $`\mathrm{\Phi }\mathrm{\Phi }\mathrm{log}N`$ , in order that $`e^\mathrm{\Phi }`$ be directly related to the โt Hooft coupling. The (independent) field equations then read
$`2\ddot{\mathrm{\Phi }}4\dot{\mathrm{\Phi }}^2+8\dot{\mathrm{\Phi }}\dot{\lambda }+10\dot{\mathrm{\Phi }}\dot{\nu }+3{\displaystyle \frac{C}{N}}e^\mathrm{\Phi }`$ $`=`$ $`0,`$ (18)
$`\ddot{\lambda }(2\dot{\mathrm{\Phi }}4\dot{\lambda }5\dot{\nu })\dot{\lambda }\frac{1}{2}e^{2\mathrm{\Phi }10\nu }+\frac{1}{4}{\displaystyle \frac{C}{N}}e^\mathrm{\Phi }`$ $`=`$ $`0,`$ (19)
$`\ddot{\nu }(2\dot{\mathrm{\Phi }}4\dot{\lambda }5\dot{\nu })\dot{\nu }4e^{2\nu }+\frac{1}{2}e^{2\mathrm{\Phi }10\nu }+\frac{1}{4}{\displaystyle \frac{C}{N}}e^\mathrm{\Phi }`$ $`=`$ $`0.`$ (20)
From these one can neatly understand the role of the dilaton tadpole: it represents a $`1/N`$ correction to the leading type IIB supergravity equations, and thus induces $`1/N`$ corrections to the $`AdS_5\times \text{โโ}^5`$ leading geometry, in agreement with field theory expectations. Moreover, $`1/N`$ is a natural small parameter in the theory and suggests to solve the equations by iterations
$`\mathrm{\Phi }(\tau )`$ $`=`$ $`\mathrm{\Phi }_0(\tau )+{\displaystyle \frac{1}{N}}\mathrm{\Phi }_1(\tau )+\mathrm{},`$ (21)
$`\lambda (\tau )`$ $`=`$ $`\lambda _0(\tau )+{\displaystyle \frac{1}{N}}\lambda _1(\tau )+\mathrm{},`$ (22)
$`\nu (\tau )`$ $`=`$ $`\nu _0(\tau )+{\displaystyle \frac{1}{N}}\nu _1(\tau )+\mathrm{}.`$ (23)
The leading order equations
$`2\ddot{\mathrm{\Phi }}_04\dot{\mathrm{\Phi }}_0^2+8\dot{\mathrm{\Phi }}_0\dot{\lambda }_0+10\dot{\mathrm{\Phi }}_0\dot{\nu }_0`$ $`=`$ $`0,`$ (24)
$`\ddot{\lambda }_0(2\dot{\mathrm{\Phi }}_04\dot{\lambda }_05\dot{\nu }_0)\dot{\lambda }_0\frac{1}{2}e^{2\mathrm{\Phi }_010\nu _0}`$ $`=`$ $`0,`$ (25)
$`\ddot{\nu }_0(2\dot{\mathrm{\Phi }}_04\dot{\lambda }_05\dot{\nu }_0)\dot{\nu }_04e^{2\nu _0}+\frac{1}{2}e^{2\mathrm{\Phi }_010\nu _0}`$ $`=`$ $`0,`$ (26)
independent of the dilaton tadpole, are consistently solved by the expected (leading) $`AdS_5\times \text{โโ}^5`$ geometry
$`\mathrm{\Phi }_0=\phi ,`$ (27)
$`\lambda _0=\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\tau ,`$ (28)
$`\nu _0=\frac{1}{8}\mathrm{log}8+\frac{1}{4}\phi .`$ (29)
The next-to-leading order equations then reduce to
$`\ddot{\mathrm{\Phi }}_1+4\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\dot{\mathrm{\Phi }}_1+\frac{3}{2}Ce^\phi =0,`$ (30)
$`\ddot{\lambda }_1+\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\left(2\dot{\mathrm{\Phi }}_1+8\dot{\lambda }_1+5\dot{\nu }_1\right)8\sqrt[4]{8}e^{{\scriptscriptstyle \frac{1}{2}}\phi }\left(\mathrm{\Phi }_15\nu _1\right)+\frac{1}{4}Ce^\phi =0,`$ (31)
$`\ddot{\nu }_1+4\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\dot{\nu }_1+8\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{2}}\phi }(\mathrm{\Phi }_15\nu _1)+\frac{1}{4}Ce^\phi =0,`$ (32)
and are solved by
$`\mathrm{\Phi }_1=\frac{3}{8\sqrt[8]{8}}Ce^{{\scriptscriptstyle \frac{5}{4}}\phi }\tau ,`$ (33)
$`\lambda _1=\frac{3}{64}Ce^\phi \tau ^2\frac{15}{256\sqrt[8]{8}}Ce^{{\scriptscriptstyle \frac{5}{4}}\phi }\tau ,`$ (34)
$`\nu _1=\frac{3}{32\sqrt[8]{8}}Ce^{{\scriptscriptstyle \frac{5}{4}}\phi }\tau \frac{1}{32}8^{{\scriptscriptstyle \frac{5}{4}}}Ce^{{\scriptscriptstyle \frac{3}{2}}\phi }.`$ (35)
In principle one should also consider the higher order iterations. However, we shall stop here since in the low energy action we are already neglecting contributions of higher order in $`1/N`$, as for example a cosmological constant.
Collecting the various contributions, one finds the following expressions for the dilaton
$$\mathrm{\Phi }=\phi \frac{3}{8}\frac{1}{\sqrt[8]{8}}\frac{C}{N}e^{\frac{5}{4}\phi }\tau ,$$
(36)
and for the metric tensor
$`{\displaystyle \frac{1}{\alpha ^{}}}ds^2`$ $`=`$ $`d\tau ^2+\mathrm{exp}\left[2\sqrt[8]{8}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\tau +\frac{3}{32}{\displaystyle \frac{C}{N}}e^\phi \tau ^2\right]\eta _{ab}dx^adx^b`$ (37)
$`+\mathrm{exp}\left[\frac{1}{4}\mathrm{log}8+\frac{1}{2}\phi \frac{3}{16\sqrt[8]{8}}{\displaystyle \frac{C}{N}}e^{{\scriptscriptstyle \frac{5}{4}}\phi }\tau \right]d\mathrm{\Omega }_5^2,`$
It is not hard to realize that the introduction of a dilaton tadpole results into a running dilaton and running radii for the $`AdS_5`$ and $`\text{โโ}^5`$ spaces. This can be better appreciated if one parametrizes the metric as
$$\frac{1}{\alpha ^{}}ds^2=R^2(u)\frac{du^2}{u^2}+\frac{u^2}{R^2(u)}\eta _{ab}dx^adx^b+\stackrel{~}{R}^2(u)d\mathrm{\Omega }_5^2,$$
(38)
in terms of the energy scale
$$u=e^{\lambda (\tau )}๐\tau ,$$
(39)
with the radii
$$R(u)=\frac{d\tau }{d\mathrm{log}u},\stackrel{~}{R}(u)=e^{\nu (\tau (u))}.$$
(40)
Collecting our previous results, we find
$`\tau `$ $`=`$ $`\frac{1}{\sqrt[8]{8}}e^{{\scriptscriptstyle \frac{1}{4}}\phi }\left(\mathrm{log}u\frac{3}{64}\frac{C}{N}e^\phi \mathrm{log}^2u\right),`$ (41)
$`R^2(u)`$ $`=`$ $`{\displaystyle \frac{{\scriptscriptstyle \frac{1}{\sqrt[4]{8}}}e^{\frac{1}{2}\phi }}{1+\frac{3}{16}\frac{C}{N}e^\phi \mathrm{log}u}},`$ (42)
$`\stackrel{~}{R}^2(u)`$ $`=`$ $`{\displaystyle \frac{{\scriptscriptstyle \frac{1}{\sqrt[4]{8}}}e^{\frac{1}{2}\phi }}{1+\frac{3}{16\sqrt[4]{8}}\frac{C}{N}e^{2\phi }\mathrm{log}u}},`$ (43)
$`e^\mathrm{\Phi }`$ $`=`$ $`e^\phi +\frac{3}{8\sqrt[4]{8}}\frac{C}{N}e^\phi \mathrm{log}u.`$ (44)
At this point, before we interpret these results, various comments are in order. First, in the above expressions we are repeatedly ignoring $`1/N^2`$ sub-leading terms. Moreover, the value of $`e^\phi `$ can not be interpreted any longer as the โt Hooft coupling, since now the field theory is not conformal and the gauge coupling runs. In addition, from eqs. (42) and (43) one might argue that $`R`$ and $`\stackrel{~}{R}`$ develop singularities at some point in the IR (notice that for our model $`C`$ is positive and then singularity presents itself for $`\mathrm{log}u<0`$). This, however, is not the case. The singularities are at points beyond the validity of the $`1/N`$ approximation and are an artifact of our choice of coordinates. Note, in fact, that the metric (37) is well defined for every energy scale. Finally, contrary to the familiar $`๐ฉ=4`$ super-conformal case the radii $`R`$ and $`\stackrel{~}{R}`$ have now different values. This is expected since at order $`1/N`$ the dilaton tadpole behaves like an effective cosmological constant, and therefore (at that order) the overall curvature of the ten-dimensional space is no longer vanishing.
We can now turn to the interpretation of eqs. (42) and (44), in the spirit of AdS/CFT correspondence. The behavior of the dilaton is consistent with the fact that, in the dual gauge theory, no longer conformal, the coupling constant is expected to run. More precisely, once the dilaton is identified with the โt Hooft coupling, $`e^{\mathrm{\Phi }(u)}\alpha (u)`$, eq. (44) reproduces precisely the logarithmic behavior (5) expected from the gauge theory side, as already shown in . Moreover, in the large $`N`$ limit the running of the coupling constant is suppressed, in agreement with the fact that in this limit the theory is conformal. Following , it would be interesting to compute also in this case the numerical value of the dilaton tadpole $`C`$ and compare it with the one-loop beta-function coefficient $`b_1=\frac{16}{3}`$. To this end, however, one should compute the contributions from closed strings in the $`AdS_5\times \text{โโ}^5`$ background with unoriented world-sheet wrapping the $`\text{โโ}^5`$. Unfortunately, computations in genuinely curved backgrounds are still an open problem in string perturbation theory.
Following the running of the AdS radius plays a natural role in the computation of the quark anti-quark potential, the subject of the next section.
## 4. The Wilson loop
We can now use the โrunningโ $`AdS_5\times \text{โโ}^5`$ geometry to compute the quark anti-quark potential using the holographic Wilson loop (see for a recent review). In it was suggested that a natural description of the Wilson loop in terms of gravity is
$$W(๐)e^S,$$
(45)
where, in the limit of large โt Hooft coupling, $`S`$ is the proper area of a string world-sheet
$$S=\frac{1}{2\pi \alpha ^{}}๐\tau ๐\sigma \sqrt{\mathrm{det}g_{\mu \nu }_\alpha X^\mu _\beta X^\nu }$$
(46)
describing the loop $`๐`$ on the boundary of AdS. The quark anti-quark potential can then be extracted choosing the standard rectangular loop with sides $`L`$ and $`T`$, with $`T\mathrm{}`$.
Using the parametrization (38) for the metric $`g_{\mu \nu }`$, the Nambu-Goto action reads
$$S/T=E=๐x\sqrt{(_xu)^2+u^4/R^4(u)},$$
(47)
and thus the quark anti-quark separation and the interaction energy are
$$L\frac{R_0^2}{u_0}\frac{dy}{y^2\sqrt{y^4R_0^4/R^4(yu_0)}},$$
(48)
and
$$Eu_0\frac{y^2dy}{\sqrt{y^4R_0^4/R^4(yu_0)}},$$
(49)
with $`R_0=R(u_0)`$.
If the energy $`u_0`$ is within the range of validity of the expansion, $`\frac{1}{N}\mathrm{log}u_01`$, the ratio $`R_0^4/R^4(yu_0)`$ can be approximated by
$$\frac{R_0^4}{R^4(yu_0)}\frac{1+{\scriptscriptstyle \frac{3}{8}}{\scriptscriptstyle \frac{C}{N}}e^\phi \mathrm{log}yu_0}{1+\frac{3}{8}\frac{C}{N}e^\phi \mathrm{log}u_0}1+\frac{3}{8}\frac{C}{N}e^\phi \mathrm{log}y.$$
(50)
Thus, one has
$$L\frac{R_0^2}{u_0},\mathrm{and}Eu_0,$$
(51)
and the potential
$$E\frac{R_0^2}{L}\frac{R^2(1/L)}{L}=\frac{\frac{1}{\sqrt[4]{8}}e^{\frac{1}{2}\phi }}{L(1\frac{3}{16}\frac{C}{N}e^\phi \mathrm{log}L)}$$
(52)
can be expressed in terms of the running AdS radius, similarly to .
Eq. (52) reproduces precisely the behavior expected from the gauge theory side: at finite $`N`$ the Coulomb phase is characterized by a quark anti-quark potential (6) with a (logarithmically) running gauge coupling. The fact that the two prescriptions, the dilaton running
$$E=\frac{\alpha (L)}{L}=\frac{e^{\mathrm{\Phi }(L)}}{L}\frac{e^\phi }{L\left(1\frac{3}{8\sqrt[4]{8}}\frac{C}{N}e^{2\phi }\mathrm{log}L\right)}$$
(53)
and the Wilson loop (52), yield similar results (in contrast to tachyonic type 0 models ) is encouraging, although different powers of $`e^\phi `$ are still involved in the two cases. Note that (53) is a perturbative prescription in contrast to (52) that should, a priori, be applicable to the non-perturbative regime.
Finally, let us describe the IR behavior of the gauge theory. To this end, it is simpler to work with the metric in the form (37). The quark anti-quark energy is then
$$E=๐x\sqrt{e^{\lambda (\tau )}(_x\tau )^2+e^{2\lambda (\tau )}},$$
(54)
and, since
$$\lambda (\tau )=\alpha \tau +\frac{1}{N}\beta \tau ^2,$$
(55)
with $`\alpha `$ and $`\beta `$ positive constants, $`e^\lambda `$ has a minimum at $`\tau _{\mathrm{min}}=\alpha N/2\beta `$, that corresponds to the IR, implying a confining long distance potential
$$Ee^{\lambda _{\mathrm{min}}}L=\sigma L.$$
(56)
The reason behind this result is rather simple to understand. For large enough separation, the string will prefer to stay at the minimum of $`\lambda `$, thus minimizing the action (54). Therefore the energy will be simply proportional to the separation on the AdS boundary. Moreover, the QCD string tension (measured in units of $`1/\alpha ^{}`$)
$$\sigma =\mathrm{exp}\lambda _{\mathrm{min}}=\mathrm{exp}\left(\frac{\alpha ^2N}{4\beta }\right)=\mathrm{exp}\left(\frac{16}{3}\sqrt[4]{8}e^{{\scriptscriptstyle \frac{3}{2}}\phi }C^1N\right)$$
(57)
is exponentially suppressed with $`N`$, in agreement with field theory analysis (9).
## 5. Comments
In conclusion, let us make a few comments about the validity of the holographic description that we have just presented. The holographic conjecture relates 4d strongly coupled gauge theories to 5d (super)gravity, and is thus valid, in principle, only within a certain range of energies. Moreover, the procedure of solving the gravitational equations by iterations further restricts the solution. First, we are taking into account only the next-to-leading contribution associated to the dilaton tadpole (a $`1/N`$ effect), and therefore we need $`N1`$ to be sure that higher order string loops be negligible. Moreover, the procedure of solving the differential equations by iterations is valid only when the expansion parameter is small. Since the expansion parameter is $`ฯต\frac{1}{N}e^{{\scriptscriptstyle \frac{3}{2}}\phi }\mathrm{log}u`$, the solution is valid only in some regions of the $`(u,N)`$ space. Note, however, that for any given energy $`u`$ and initial AdS radius $`e^{{\scriptscriptstyle \frac{1}{4}}\phi }`$, it is always possible to find an $`N`$ such that $`ฯต`$ is small, compatibly with the previous requirement. Finally, the curvature of the space should be small enough to allow one to trust classical gravity and to neglect $`\alpha ^{}`$ corrections. Once the previous requirements are met, this last constraint simply translates into large $`e^{{\scriptscriptstyle \frac{1}{4}}\phi }`$, as in the $`๐ฉ=4`$case.
Let us now see whether our solution meets these requirements. It is not hard to see that large $`e^\phi `$ and small expansion parameter $`ฯต`$ translate into large โt Hooft coupling. But, from the field theory considerations, one expects the logarithmic behavior for small and intermediate coupling (5). One should assume non-large $`e^\phi `$ and therefore $`\alpha ^{}`$ corrections should not be neglected, thus invalidating our solution. However, this is not the case. A more careful analysis reveals that our results depend only on the existence of an $`AdS_5\times \text{โโ}^5`$ solution at infinite $`N`$. Now, since it is a common lore that $`\alpha ^{}`$ corrections do not change the geometry of the space (namely, that $`๐ฉ=4`$is described by an $`AdS_5\times S^5`$ metric at any coupling), the logarithmic running should not be affected. Moreover, there seem to be problems in the computation of the short-distance Wilson loop. First, the direct Wilson loop computation results into a value for the one-loop beta-function different from that obtained by the running dilaton recipe. This is actually related to $`\alpha ^{}`$ corrections that, as in the $`๐ฉ=4`$case , are expected to affect the parameters, but not the qualitative picture. In addition, the string world-sheet โinvadesโ also regions where $`ฯต`$ is not small. Namely, the holographic computation of the Wilson loop for a given quark anti-quark separation $`L`$ requires that the metric be known everywhere throughout the range $`(u1/L,\mathrm{})`$. In order to avoid this problem and thus be able to trust our solution, the world-sheet should extend deeply inside AdS, and thus the potential (52) applies only to not too small quark anti-quark separations (i.e. our results are not valid in the extreme UV). Finally, the confining phase for long distances relies on the existence of a minimum for $`\lambda (\tau )`$. However, at $`\tau =\tau _{\mathrm{min}}`$ the contribution of next-to-leading term $`\lambda _1`$ is of the same order as $`\lambda _0`$. Thus, we can not exclude that higher order contributions could eliminate the minimum, but hopefully the only effect of higher corrections could be simply to shift $`\tau _{\mathrm{min}}`$. However, a definite answer to this problem requires an exact solution to the exact low-energy action.
Acknowledgments. We are grateful to E. Gardi, A. Sagnotti, D. Seminara and especially to G. Grunberg and J. Sonnenschein for useful comments. A.A. thanks the Department of Applied Mathematics and Theoretical Physics at Cambridge University for the warm hospitality while this work was being completed. This research was supported in part by EEC under TMR contract ERBFMRX-CT96-0090. |
warning/0003/cond-mat0003201.html | ar5iv | text | # Electronic Fine Structure in the Electron-Hole Plasma in SrB6
\[
## Abstract
Electron-hole mixing-induced fine structure in alkaline earth hexaborides leads to lower energy (temperature) scales, and thus stronger tendency toward an excitonic instability, than in their doped counterparts (viz. Ca<sub>1-x</sub>La<sub>x</sub>B<sub>6</sub>, $`x0.005)`$, which are high Curie temperature, small moment ferromagnets. Comparison of Fermi surfaces and spectral distributions with de Haas - van Alphen (dHvA), optical, transport, and tunneling data indicates that SrB<sub>6</sub> remains a fermionic semimetal down to (at least) 5 K, rather than forming an excitonic condensate. For the doped system the Curie temperature is higher than the degeneracy temperature.
\]
The observation of very low moment magnetism (0.07 $`\mu _B`$ per carrier) in the very low carrier density doped semimetal Ca<sub>0.995</sub>La<sub>0.005</sub>B<sub>6</sub> (one carrier per $`6\times 6\times 6`$ unit cells) at remarkably high temperature (T<sub>c</sub> = 600 K) by Young et al. has fueled rapid investigation into possible mechanisms for this novel occurrence. The focus has been on some sort of excitonic instability, possibly an excitonic condensate that breaks time reversal symmetry and thereby allows ferromagnetism. What has not yet drawn the same attention is the undoped system, which is really the more suitable candidate for an excitonic instability than the doped system. The DB<sub>6</sub> hexaborides (D=Ca, Sr, Ba) would be semiconductors with a band gap of 2 - 3 eV between bonding and antibonding B$`{}_{}{}^{2}{}_{6}{}^{}`$ $`2p`$ bands except for a D atom $`d`$ band that dives through the gap from above, with a minimum at the X point. The result is a very small band overlap $``$90 meV at the three Brillouin zone (BZ) edge points X.
The study of instabilities in such semimetals goes back to Keldysh and Kopaev and des Cloizeaux. In the presence of a residual screened Coulomb interaction between the carriers and sufficient nesting, there are possibilities of charge density wave (CDW) or spin density wave (SDW) (โsingletโ or โtripletโ) instabilities depending on the degree of nesting. Zhitomirsky et al. have suggested that magnetism can appear as a result of consecutive CDW and SDW instabilities in a doped excitonic insulator, with polarization becoming allowed because both inversion and time-reversal symmetries are broken. Barzykin and Gorโkov contend that this model of nested semimetals is instead unstable to the appearance of a superstructure. Balents and Varma have considered the relative stability of the various order parameters involving charge, spin, and X-point pockets and the dependence on doping concentration $`x`$, and conclude that intrapocket condensation is favored over interpocket pairing.
Each of these discussions presumes the instability of the undoped system, a question that has been revisited by Zhitomirsky and Rice. They emphasize the importance of interpocket scattering processes and reemphasize that the undoped hexaborides should be good candidates for a condensed excitonic state. What we show in this paper, by comparison with band structure derived quantities with the considerable data available for SrB<sub>6</sub> , is evidence that SrB<sub>6</sub> remains a fermionic semimetal as described by conventional band theory down to at least 5 K, and perhaps down to 0.5 K where a transition to a more conductive (not insulating) phase has been observed. The electronic structure of the $`x`$=0.005 system is in fact much simpler (in the absence of an excitonic instability) and the doping level and observed Fermi surface volume is consistent with a fully polarized ferromagnet (FM).
The excitonic instability is entirely dependent on the band structure. Some of the band characteristics have been presented by Hasegawa and Yanase and by Massidda et al. and crystal stability has been studied by Ripplinger et al. Here we address the band structure in more detail than heretofore, and find that the calculated energy scales are in excellent agreement with the variety of data on SrB<sub>6</sub>. The fine structure that we discuss applies to the undoped systems and affects the tendency toward instability, and it can be tested with dHvA data if the system remains fermionic at low temperature (T). We have calculated the band structures of these compounds using the accurate, full potential augmented plane wave method. In these hexaborides the differences between the local density approximation and the generalized gradient approximation (GGA) are small even on the scale of the fine structure we will be discussing, and we use the GGA results. We discuss primarily SrB<sub>6</sub> because of the more extensive experimental data, and because when doped, SrB<sub>6</sub> also shows (as does BaB<sub>6</sub> ) a FM phase similar to doped CaB<sub>6</sub>, but at even higher temperature. We use the measured lattice constant of 4.20 ร
and internal parameter (B position along the cubic axis) of 0.203, the latter confirmed by our total energy minimization. We have checked that spin-orbit coupling has no noticable effect on the results we discuss.
The two bands that overlap at the X=(100)$`\pi /a`$ point, shown in Fig. 1, are characterized at the most basic level by longitudinal ($`m_l`$) and transverse ($`m_t`$) masses, given by $`m_l`$=0.50, -2.13 and $`m_t`$=0.21, -0.20 relative to the free electron mass (hole masses are negative). Thermal masses ($`m_t^2m_l)^{1/3}`$ are 0.28 and -0.44 respectively. The band overlap (โnegative gapโ) is 90 meV in SrB<sub>6</sub> . Crossing of the electron ($`e`$) and hole ($`h`$) bands only occurs precisely along the $`\mathrm{\Delta }`$ = ($`\xi `$,0,0)$`\frac{\pi }{a}`$ and Z = (1,$`\xi `$,0)$`\frac{\pi }{a}`$ lines; elsewhere coupling results in anticrossing. This mixing results in the two overlapping and intersecting ellipsoids (in the absence of coupling) becoming two distinct surfaces as shown in Fig. 2: a circular lens centered on $`\mathrm{\Delta }`$ containing holes, and a ring (โnapkin ringโ) centered at X containing electrons. Each surface is made of pieces of each of the two ellipsoids, and each contains an $`e`$ and a $`h`$ part.
We have used the condition that the number of electrons must equal the number of holes to determine the intrinsic Fermi energy accurately. The procedure becomes quite delicate, requiring energy grids of spacing $`\frac{1}{150}`$ $`\frac{\pi }{a}`$ to represent the hybridization sufficiently accurately. The result is that, whereas the ellipsoids at equal volumes would contain $`n_o`$=2$`\times 10^3`$ carriers, the six lenses and three rings contain only $`n_e/2=n_h/2`$ = 2.7$`\times 10^4`$ of the BZ volume.
The effect of band mixing is most easily seen in the drastic effect it has on the density of states (DOS). As pictured in Fig. 3, instead of simple overlapping $`(EE_o)^{1/2}`$ edges, there are rather strong peaks near the points of mixing (due to flattened anticrossing bands) with a deep minimum almost at the $`x`$=0 Fermi level. This structure makes the DOS at the Fermi level, N(E<sub>F</sub>), depend much more strongly on carrier concentration than might have been guessed.
In free space this plasma ($`n_e=n_h`$) would be characterized by an electron gas parameter $`r_s`$ = 60, which would be an extremely low density plasma. In a solid the effective mass (on average $`|m^{}|0.35`$, see above) and the background dielectric constant $`ฯต`$ rescale the kinetic and potential energies, respectively, and alter the effective density. The value of $`ฯต`$ has not been reported. Judging from the fact that SrB<sub>6</sub> would have an average direct gap of around 3 eV except for the single conduction band that dips down at X, and that this gap is similar to that of Si ($`ฯต`$=12) but the crystal is less covalent and partially ionic, we estimate $`ฯต8`$ for the background dielectric constant. This value leads to plasma parameters
$`a_B^{}`$ $`=`$ $`a_Bฯต{\displaystyle \frac{m}{m^{}}}24a_B13\AA ,`$ (1)
$`r_s^{}`$ $`=`$ $`r_s{\displaystyle \frac{a_B}{a_B^{}}}2.5`$ (2)
$`E_R^{}`$ $`=`$ $`E_R{\displaystyle \frac{m^{}}{m}}{\displaystyle \frac{1}{ฯต^2}}25meV,`$ (3)
here $`a_B`$ is the Bohr radius. The Rydberg E$`{}_{}{}^{}{}_{R}{}^{}`$, is the exciton binding energy that sets the scale of bandgaps for which a low temperature excitonic instability is possible. Using these values, the $`e`$ and $`h`$ densities of SrB<sub>6</sub> falls within the range of effective density of alkali metals, but with an excitonic energy scale of E$`{}_{R}{}^{}`$ 25 meV that is two orders of magnitude smaller than in alkali metals. The carrier plasma energy is changed little by this renormalization \]$`(ฯต^1m/m^{})^{1/2}1/2`$\] and remains on the scale of 100 meV. The corresponding Fermi energies that determine the degeneracy temperature are 20 meV for the holes and 15 meV for the electrons, making the degeneracy temperature T$`{}_{F}{}^{}`$250 K.
Barring an excitonic instability, at low temperature T$`<<`$T<sub>F</sub> dHvA oscillations will measure the cross sectional areas and cyclotron masses of the extremal Fermi surface orbits perpendicular to the applied field. For the field along a cubic axis, these Fermi surfaces lead to five extremal orbits whose band masses we have calculated from the energy derivative of the cross sectional area $`m^{}=\frac{\mathrm{}^2}{2\pi }dA/dE`$. The lens and ring cross sections $`O_{lens,0}`$ and $`O_{ring,0}`$ lie in the plane of Fig. 2. There are three extremal orbits perpendicular to these (circling the $`\mathrm{\Delta }`$ line): $`O_{lens,1}`$, $`O_{ring,in}`$, and $`O_{ring,out}`$, where $`in`$ and $`out`$ denote the inner and outer ring orbits.
The areas and cyclotron band masses are given in Table I. The values of the three observed frequencies are 30, 48, and 308 Tesla. The first two are in the range of our values, and in fact agree well with the two lens orbits. These two orbits, both areas and masses, arise from the mixing that eliminates the intersections of ellipsoids and creates new orbits that can be seen in Fig. 2. The 308 Tesla orbit is rather large to be accounted for in terms of band structure results; even with magnetic breakdown, the largest orbit (for the cigar ellipsoid) is only 196 Tesla. It is interesting that the available data is consistent with โmissingโ rings, but it is also possible that ring orbits are simply more difficult to observe.
The distinctive characters of the carriers may play an important role in the behavior of the hexaborides, e.g. by affecting matrix elements or, due to La doping on the alkaline earth sublattice, reducing the mean free path for the electrons much more than for the holes. The band that dips down through the gap has Sr $`4d`$ character (Fig. 3), and this band forms the electron ellipsoids. Thus the $`e`$ carriers are a combination of Sr $`d`$ and B $`p`$, while the $`h`$ carriers are purely bonding B $`p`$ character. Due to the mixing that rehybridizes the bands and leads to the lenses and rings, each Fermi surface is roughly half $`e`$-like and half $`h`$-like, and thus partly Sr $`d`$ as well as B $`p`$.
We now compare our results with available data on SrB<sub>6</sub> . The band structures indicate a degeneracy scale E of 15-20 meV (degeneracy temperature T$`{}_{}{}^{}`$ 250 K), reduced from 600 K by anticrossing electron and hole bands. Interband transitions in the infrared should peak at the separation of DOS peaks in Fig. 3, near $`\omega ^{}`$ 40 meV/$`\mathrm{}`$ 320 cm<sup>-1</sup>. The experimental data indicate: (1) a change from metallic slope of resistivity $`\rho `$ to non-metallic slope at 250 K $``$ T, consistent with some change as the system evolves from non-degenerate to degenerate (the transport is not understood), (2) a rise in the optical conductivity (at low T) below 300 cm$`{}_{}{}^{1}\omega ^{}`$, but the peaking at 40-80 cm<sup>-1</sup> (i.e. below 10 meV) may reflect k-conserving processes not identifiable in the DOS of Fig. 3, (3) a broadening and weakening of this absorption peak on the scale of 200-300 K $``$ T, (4) a strong onset of interband transitions at $``$2 eV, roughly consistent with the band structure, (5) negligible electronic contribution to the specific heat above 5 K; the band structure value would not be detectable, (6) tunneling spectra show a conductance minimum at a negative bias of -40 meV $`\mathrm{}\omega ^{}`$, and strong T dependence of the spectrum at -40 meV and at zero bias. All of these observations are consistent, almost quantitatively, with the band overlap and mixing that we calculate.
There are additional observations: $`\rho (T)`$ shows structure at 50-60 K, and drops very rapidly by 30% below 0.5 K, this latter temperature corresponding to an energy scale of 40 $`\mu `$eV for which there is no immediate explanation. There is excess specific heat (above the lattice contribution) below 5 K, but with only a small integrated entropy of less than 0.05% of $`Rln2`$. Except for this low temperature regime, observations seem consistent with a fermionic plasma with band structure as described above.
Now we turn to the doped magnetic system. Doping of $`x`$=0.005 increases the Fermi level by 50 meV as shown in Fig. 1, for which there is a single ellipsoidal Fermi surface (at each of the X points) with eccentricity $`\eta `$= k/k = 1.6. The hole surface vanishes near $`x`$=0.004. For doped ($`x`$=0.005) CaB<sub>6</sub>, the reported dHvA frequencies of 350 T and 495 T are consistent with such ellipsoids with eccentricity $`\eta `$=495/350=1.4. The required values of k and k correspond to a doping level of 0.0090 for a paramagnetic system; however, if the ellipsoids represent a fully polarized electron gas, the corresponding doping level is 0.0045, very close to the nominal doping level. The reported ordered moment of 0.07 $`\mu _B`$ per carrier might reflect large magnetic fluctuations, noncollinearity of spins, orbital currents, or phase separation in the magnetic phase.
The origin of the magnetism has had no explanation except as an outgrowth of an excitonic undoped phase, which phase in SrB<sub>6</sub> seems not to be supported by the data. At $`x`$=0.005 and higher, the electronic structure of the paramagnetic systems seems to be very simple. Inter-pocket $`ee`$ nesting between inequivalent X points is the only $`\omega 0`$ nesting (Q$`0)`$, and is maximum between points on their equators. If the Fermi surfaces were spheres, $`\omega `$ = 0 nesting would be perfect for Q=(1,1,0)$`\pi /a`$ wavevectors, tending to drive CDW or SDW instabilities. Due to the eccentricity (i.e. $`m_tm_l`$), these scattering processes are smeared by $`\delta `$Q$`2\times 10^2\pi /a`$ for $`\omega `$=0 or by $`\delta \omega 30`$ meV at Q itself. Moreover, the system is already magnetic above the degeneracy temperature, where nesting is an irrelevant process. At high T a nonzero โ but rather small โ density of holes will be thermally excited, leading perhaps to a charge-unbalanced, quasiclassical $`eh`$ plasma, but the high temperature should also preclude formation of an exciton-fermion plasma.
To summarize, we have looked closely at the electronic structure of undoped and doped divalent hexaborides, and compared them with the available transport, thermal, optical, dHvA, and tunneling data. The behavior of SrB<sub>6</sub> above 5 K is consistent with expectations, and energy scales, obtained from the band structure. The observed Fermi surface of $`x`$=0.005 doped CaB<sub>6</sub> is most easily accounted for as the majority Fermi surface of a fully polarized electron gas, but this interpretation is not readily reconciled with the measured tiny moment.
We acknowledge helpful communications with Z. Fisk, R. G. Goodrich, and J. W. Allen during the course of this work, and comments on the manuscript by R. Monnier. This work was supported by an NSF-CONICET International Collaboration. W. E. P. was supported by National Science Foundation (NSF) Grant No. DMR-9802076. Ruben Weht acknowledges support from Fundaciรณn Antorchas Grants No. A-13622/1-103 and A-13661/1-27. Much of this work was done when the authors were at the Institute of Theoretical Physics in Santa Barbara, which is supported by NSF Grant No. PHY 94-07194. |
warning/0003/hep-th0003086.html | ar5iv | text | # Contents
## 1 Introduction
In recent years there have been many exciting advances in our understanding of M-theory โ our best candidate for the fundamental theory of everything. The theory claims to describe physics appropriately in regions of space with high energies and large curvature scales. As these characteristics are exactly those found in the initial conditions of the universe it is only natural to incorporate M-theory into models of early universe cosmology.
The necessity to search for alternatives to the Standard Big-Bang (SBB) model of cosmology stems from a number of detrimental problems such as the horizon, flatness, structure formation and cosmological constant problems. Although inflationary models have managed to address many of these issues, inflation, at least in its current formulation, does not explain everything. In particular, inflation fails to address the fluctuation, super-Planck scale physics, initial singularity and cosmological constant problems as discussed in .
At the initial singularity, physical invariants such as the Ricci scalar, $`R`$, blow up. Other measurable quantities, for example temperature and energy density also become infinite. From the Hawking-Penrose singularity theorems we know that such spacetimes are geodesically incomplete. So, when we ask the question of how the universe began, the inevitable and unsatisfactory answer is that we donโt know. The physics required to understand this epoch of the early universe is necessarily rooted in a theory of quantum gravity. Presently, string theory is the only candidate for such a unifying theory. It is therefore logical to study the ways in which it changes our picture of cosmology. Although an ambitious aspiration, we hope that M-theory will solve the above mentioned dilemmas and provide us with a complete description of the evolution of the universe.
In this analysis, we must proceed with caution. Our present understanding of M-theory is extremely limited, as is our understanding of cosmology before the first $`10^{43}`$ seconds. Nevertheless, it is clear that the study of string cosmology is essential to the development of string theory, and extremely important for our understanding of the early universe.
The purpose of this article is to introduce some of the most promising work and themes under investigation in string cosmology. We begin with a brief, qualitative introduction to M-theory in Section 2.
In Section 3 we review the work of Brandenberger and Vafa in which the 1980s version of string theory is used to solve the initial singularity problem and in an attempt to explain why we live in four macroscopic dimensions despite the fact that string theory seems to predict the wrong number of dimensions, namely ten. We then explain how this scenario has been updated in order to include the effects of $`p`$-branes .
Section 4 provides a brief introduction to the Pre-Big-Bang scenario -. This is a theory based on the low energy effective action for string theory, developed in the early 1990s by Gasperini and Veneziano.
Another promising attempt to combine M-theory with cosmology, that of Lukas, Ovrut and Waldram , is presented in Section 5. Their work is based on the model of heterotic M-theory constructed by Hoลava and Witten and is inspired by eleven dimensional supergravity, the low energy limit of M-theory. The motivation for this work was to construct a toy cosmological model from the most fundamental theory we know.
The final section (6) reviews some models involving large extra dimensions. This section begins with a short introduction to the hierarchy problem of standard model particle physics and explains how it may be solved using large extra dimensions. โBrane Worldโ scenarios are then discussed focusing primarily on the models of Randall and Sundrum , where our four dimensional universe emerges as the world volume of a three brane. The cosmologies of such theories are reviewed, and we briefly comment on their incorporation into supergravity models, string theory and the AdS/CFT correspondence.
The sections in this review are presented more or less chronologically.<sup>2</sup><sup>2</sup>2This review is in no way comprehensive. As it is impossible to discuss all aspects of string cosmology I have included a large list of references at the end. Some of the topics I will not cover in the text may be found there. For discussions of p-brane dynamics and cosmology see -, . For recent reviews on other cosmological aspects of M-theory see . For some ideas on radically new cosmologies from M-theory see e.g. -.
## 2 M-Theory
For several years now, we have known that there are five consistent formulations of superstring theory. The five theories are ten-dimensional, two having $`N=2`$ supersymmetry known as Type IIA and Type IIB and three having $`N=1`$ supersymmetry, Type I, $`SO(32)`$ heterotic and $`E_8\times E_8`$ heterotic. Recently, duality symmetries between the various theories have been discovered, leading to the conjecture that they all represent different corners of a large, multidimensional moduli space of a unified theory named, M-theory. Using dualities we have discovered that there is a sixth branch to the M-theory moduli space (see Fig. (2)) corresponding to eleven-dimensional supergravity .
> Figure 2: This is a slice of the eleven-dimensional moduli space of M-theory. Depicted are the five ten-dimensional string theories and eleven-dimensional supergravity, which is identified with the low energy limit of M-theory.
It is possible that using these six cusps of the moduli space we have already identified the fundamental degrees of freedom of the entire nonperturbative M-theory, but that their full significance has yet to be appreciated. A complete understanding and consistent formulation of M-theory is the ultimate challenge for string theorists today and will take physicists into the new M(illennium).
## 3 Superstrings and Spacetime Dimensionality
Perhaps the greatest embarrassment of string theory is the dimensionality problem. We perceive our universe to be four dimensional, yet string theory seems to naively predict the wrong number of dimensions, namely ten. The typical resolution to this apparent conflict is to say that six of the dimensions are curled up on a Planckian sized manifold. The following question naturally arises, why is there a six/four dimensional split between the small/large dimensions? Why not four/six, or seven/three? Although there is still no official answer to this question, a possible explanation emerges from cosmology and the work of Brandenberger and Vafa which we will summarize in this section. We will then show how it is possible to generalize this scenario of the 1980s to incorporate our current understanding of string theory .
### 3.1 Duality
Before diving into the specifics of the BV model we review some basics of string dualities and thermodynamics. Consider the dynamics of strings moving in a nine-dimensional box with sides of length $`R`$. We impose periodic boundary conditions for both bosonic and fermionic degrees of freedom, so we are effectively considering string propagation in a torus. What types of objects are in our box? For one, there are oscillatory modes corresponding to vibrating stationary strings. Then, there are momentum modes which are strings moving in the box with fourier mode $`n`$ and momentum
$$p=n/R.$$
(3.1)
There are also winding modes which are strings that stretch across the box (wrapped around the torus) with energy given by
$$\omega =mR,$$
(3.2)
where $`m`$ is the number of times the string winds around the torus.
We now make the remarkable observation, that the spectrum of this system remains unchanged under the substitution
$$R\frac{1}{R},$$
(3.3)
(provided we switch the roles of $`m`$ and $`n`$). This symmetry is known as T-duality and is a symmetry of the entire M-theory, not just the spectrum of this particular model. T-duality leads us to the startling conclusion that any physical process in a box of radius $`R`$ is equivalent to a dual physical process in a box of radius $`1/R`$. In other words, one can show that scattering amplitudes for dual processes are equal. Hence, we have discovered that distance, which is an invariant concept in general relativity (GR), is not an invariant concept in string theory. In fact, we will see that many invariant notions in GR are not invariant notions in string theory. These deviations from GR are especially noticeable for small distance scales where the Fourier modes of strings become heavier (3.1) and less energetically favorable, while the winding modes become light (3.2) and are therefore more easy to create.
### 3.2 Thermodynamics of Strings
Before discussing applications of t-duality to cosmology let us review a few useful calculations of string thermodynamics. The primary assumption we will make for the following discussion is that the string coupling is sufficiently small so that we may ignore the gravitational back reaction of thermodynamical string condensates on the spacetime geometry.
String thermodynamics predicts the existence of a maximum temperature known as the Hagedorn temperature ($`T_H`$) above which the canonical ensemble approach to thermodynamics breaks down . This is due to the divergence of the partition function because of string states which exponentially increase as
$$d(E)E^p\mathrm{exp}(\beta _HE),$$
(3.4)
where $`p>0`$. The partition function is easily calculated,
$$Z=\underset{i}{}\mathrm{exp}(\beta E_i),$$
(3.5)
which diverges for $`\beta <\beta _H`$, or $`T>T_H`$.<sup>3</sup><sup>3</sup>3For more on string thermodynamics see e.g. -.
### 3.3 The BV Mechanism and the Early Universe
Consider the following toy model of a superstring-filled early universe. Besides the assumption of small coupling stated in section 3.2, we also assume that the evolution of the universe is adiabatic and make some assumptions about the size and shape of the universe.
Before the work of Brandenberger and Vafa, it was typical to speak about the process of โspontaneous compactificationโ of six of the ten dimensions predicted by string theory in order to successfully explain the origins of a large, $`3+1`$ dimensional universe. Brandenberger and Vafa proposed that, from a cosmological perspective, it is much more logical to consider the decompactification of three of the spatial directions. In other words, one starts in a universe with nine dimensions, each compactified close to the Planck length and then, for one reason or another, three spatial dimensions grow large.
The toy model of the early universe considered here is a nine dimensional box with each dimension having equal length, $`R`$. The box is filled with strings and periodic boundary conditions are imposed as described in Section (3.1).
In the SBB model it is possible to plot the scale factor $`R`$ vs. $`t`$ using the Einstein equations (Fig. (3.3)(a)). For the radiation dominated epoch, $`Rt^{1/2}`$. Furthermore, it is possible to plot $`R`$ vs. the temperature $`T`$, where $`T1/R`$ (Fig. (3.3) (b) and (c)). In string theory we have no analogue of Einsteinโs equations and hence we cannot obtain a plot of the scale factor, $`R`$ vs. $`t`$. On the other hand, we do know the entire spectrum of string states and so we can obtain an analogue of the $`R`$ vs. $`T`$ curve (see Fig. (3.3)(d)). Note that the region of Fig. (3.3)(d) near the Hagedorn temperature is not well understood, and canonical ensemble approaches break down. Fortunately, the regions to the left and right of $`T_H`$ are connected via dualities. The interested reader should see e.g. - for more modern investigations of the Hagedorn transition.
Recall, that in General Relativity the temperature $`T`$ goes to infinity as the radius $`R`$ decreases. As we have already mentioned, string theory predicts a maximum temperature, $`T_H`$ and therefore one should expect the stringy $`R`$ vs. $`T`$ curve to be drastically altered. Furthermore, we found that string theory enjoys the $`R1/R`$ symmetry which leads to a $`\mathrm{ln}R\mathrm{ln}R`$ symmetry in Fig. (3.3)(d). For large values of $`R`$, $`R1/T`$ is valid since the winding modes are irrelevant and the theory looks like a point particle theory. For small $`R`$ the $`TR`$ curve begins to flatten out, approach the Hagedorn temperature and then as we continue to go to smaller values of $`R`$ the temperature begins to decrease. This behavior is a consequence of the T-duality of string theory. As $`R`$ shrinks, the winding modes which are absent in point particle theories become lighter and lighter, and are therefore easier to produce. Eventually, (with entropy constant) the thermal bath will consist mostly of winding modes, which explains the decrease in temperature once one continues past $`T_H`$ to smaller values of $`R`$.
> Figure 3.3: In (a) (and (b)), we have plotted $`R`$ vs. $`t`$ ($`T`$ vs. $`R`$) for the SBB model. Figures (c) and (d) are plots of $`T`$ vs. $`\mathrm{ln}R`$ for both the SBB and String cosmological models respectively. Note the $`\mathrm{ln}R\mathrm{ln}R`$ symmetry in (d).
An observer traveling from large $`R`$ to small $`R`$, actually sees the radius contracting to $`R=1`$ (in Planck units) and then expanding again. This makes us more comfortable with the idea of the temperature beginning to decrease after $`R=1`$. The reason for this behavior is that the observer must modify the measuring apparatus to measure distance in terms of light states. The details for making this change of variables are described in .
Hence, the observer described above encounters an oscillation of the universe. This encourages one to search for cosmological solutions in string theory where the universe oscillates from small to large, eliminating the initial and final singularities found in (SBB) models.
### 3.4 The Dimensionality Problem
We are now ready to ask the question, how can superstring theory, a theory consistently formulated in ten dimensions give rise to a universe with only four macroscopic dimensions? This is equivalent within the context of our toy model to asking why should three of the nine spatial dimensions of our box โwantโ to expand? To address this question, note the following observation: winding modes lead to negative pressure in the thermal bath. To understand this, recall that as the volume of the box increases, the energy in the winding modes also increases (3.2). Thus the phase space available to the winding modes decreases, which brings us to the conclusion that winding modes would โlikeโ to prevent expansion. The point is that it costs a lot of energy to expand with winding modes around. Thermal equilibrium demands that the number of winding modes must decrease as $`R`$ increases (since the winding modes become heavier). Therefore, we conclude that expansion can only occur when the system is in thermal equilibrium, which favors fewer of the winding states as $`R`$ increases. If, on the other hand, the winding modes are not in thermal equilibrium they will become plentiful and thus any expansion will be slowed and eventually brought to a halt.
Thermal equilibrium of the winding modes requires string interactions of the form
$$W+\overline{W}unwoundstates.$$
(3.6)
Here $`W`$ is a winding state and $`\overline{W}`$ is a winding state with opposite winding as depicted in Fig. (3.4).
> Figure 3.4: Strings that interact with opposite windings become unwound states.
In order for such processes to occur, the strings must come to within a Planck length of one another. As the winding strings move through spacetime they span out two dimensional world sheets. In order to interact, their worldsheets must intersect, but in a nine dimensional box the strings will probably not intersect because $`2+2<9+1`$. Since there is so much room in the box, the strings will have a hard time finding one another in order for their worldsheets to intersect and therefore it is unlikely that they will unwind. If the winding strings do not unwind, and the box starts to expand, the winding states will fall out of thermal equilibrium and the expansion will be halted.
The conclusion is that the largest spacetime dimensionality consistent with maintaining thermal equilibrium is four. Since, $`2+2=3+1`$, and therefore the largest number of spatial dimensions which can expand is three. In the next section we will see how this scenario can be incorporated into our current understanding of string theory.
### 3.5 Brane Gases and the ABE Mechanism
Recent developments in M/string theory have revealed that strings are not the only fundamental degrees of freedom in the theory. The spectrum of fundamental states also includes higher dimensional extended objects known as D-branes. Here we will examine the way in which the BV scenario unfolds in the presence of D-branes in the early universe as constructed by Alexander, Brandenberger and Easson (ABE) . Specifically, we are interested in finding out if the inclusion of branes affects the cosmological implications of . Note that this approach to string cosmology is in close analogy with the starting point of the standard big-bang model and is very different from other cosmological models which have attempted to include D-branes, for example the brane-world scenarios discussed in Sections 5 and 6. However, possible relations between this model and brane-world scenarios will be discussed later.
Our initial state will be similar to that of . We assume that the universe started out close to the Planck length, dense and hot and with all degrees of freedom in thermal equilibrium. As in , we choose a toroidal geometry in all spatial dimensions. The initial state will be a gas composed of the fundamental branes in the theory. We will consider 11-dimensional M-theory compactified on $`S^1`$ to yield 10-dimensional Type II-A string theory. The low-energy effective theory is supersymmetrized dilaton gravity. Since M-theory admits the graviton, 2-branes and 5-branes as fundamental degrees of freedom, upon the $`S^1`$ compactification we obtain 0-branes, strings (1-branes), 2-branes, 4-branes, 5-branes, 6-branes and 8-branes in the 10-dimensional universe.
The details of the compactification will not be discussed here, however we will briefly mention the origins of the above objects from the fundamental eleven-dimensional, M-theory perspective. The 0-branes of the II-A theory are the BPS states of nonvanishing $`p_{10}`$. In M-theory these are the states of the massless graviton multiplet. The 1-brane of the II-A theory is the fundamental II-A string which is obtained by wrapping the M-theory supermembrane around the $`S_1`$. The 2-brane is just the transverse M2-brane. The 4-branes are wrapped M5-branes. The 5-brane of the II-A theory is a solution carrying magnetic NS-NS charge and is an M5-brane that is transverse to the eleventh dimension. The 6-brane field strength is dual to that of the 0-brane, and is a KK magnetic monopole. The 8-brane is a source for the dilaton field .
The low-energy bulk effective action for the above setup is
$`S_{bulk}={\displaystyle \frac{1}{2\kappa ^2}}{\displaystyle }d^{10}x\sqrt{G}e^{2\varphi }[R`$ $`+`$ $`4G^{\mu \nu }_\mu \varphi _\nu \varphi `$ (3.7)
$``$ $`{\displaystyle \frac{1}{12}}H_{\mu \nu \alpha }H^{\mu \nu \alpha }],`$
where $`G`$ is the determinant of the background metric $`G_{\mu \nu }`$, $`\varphi `$ is the dilaton, $`H`$ denotes the field strength corresponding to the bulk antisymmetric tensor field $`B_{\mu \nu }`$, and $`\kappa `$ is determined by the 10-dimensional Newton constant in the usual way.
For an individual $`p`$-brane the action is of the Dirac-Born-Infeld form
$$S_p=T_pd^{p+1}\zeta e^\varphi \sqrt{det(g_{mn}+b_{mn}+2\pi \alpha ^{}F_{mn})}$$
(3.8)
where $`T_p`$ is the tension of the brane, $`g_{mn}`$ is the induced metric on the brane, $`b_{mn}`$ is the induced antisymmetric tensor field, and $`F_{mn}`$ the field strength tensor of gauge fields $`A_m`$ living on the brane. The total action is the sum of the bulk action (3.7) and the sum of all of the brane actions (3.8), each coupled as a delta function source (a delta function in the directions transverse to the brane) to the 10-dimensional action.
In the string frame the tension of a $`p`$-brane is
$$T_p=\frac{\pi }{g_s}(4\pi ^2\alpha ^{})^{(p+1)/2},$$
(3.9)
where $`\alpha ^{}l_{st}^2`$ is given by the string length scale $`l_{st}`$ and $`g_s`$ is the string coupling constant.
In order to discuss the dynamics of this system, we will need to compute the equation of state for the brane gases for various $`p`$. There are three types of modes that we will need to consider. First, there are the winding modes. The background space is $`T^9`$, and hence a $`p`$-brane can wrap around any set of $`p`$ toroidal directions. These modes are related by t-duality to the momentum modes corresponding to center of mass motion of the branes. Finally, the modes corresponding to fluctuations of the branes in the transverse directions are (in the low-energy limit) described by scalar fields on the brane, $`\varphi _i`$. There are also bulk matter fields and brane matter fields.
We are mainly interested in the effects of winding modes and transverse fluctuations to the evolution of the universe, and therefore we will neglect the antisymmetric tensor field $`B_{\mu \nu }`$. We will take our background metric with conformal time $`\eta `$ to be
$$G_{\mu \nu }=a(\eta )^2diag(1,1,\mathrm{},1),$$
(3.10)
where $`a(\eta )`$ is the cosmological scale factor.
If the transverse fluctuations of the brane and the gauge fields on the brane are small, the brane action can be expanded as
$`S_{brane}`$ $`=`$ $`T_p{\displaystyle d^{p+1}\zeta a(\eta )^{p+1}e^\varphi }`$
$`e^{\frac{1}{2}trlog(1+_m\varphi _i_n\varphi _i+a(\eta )^22\pi \alpha ^{}F_{mn})}`$
$`=`$ $`T_p{\displaystyle d^{p+1}\zeta a(\eta )^{p+1}e^\varphi }`$
$`(1+{\displaystyle \frac{1}{2}}(_m\varphi _i)^2\pi ^2\alpha _{}^{}{}_{}{}^{2}a^4F_{mn}F^{mn}).`$
The first term in the parentheses in the last line represents the brane winding modes, the second term corresponds to the transverse fluctuations, and the third term to brane matter. In the low-energy limit, the transverse fluctuations of the brane are described by a free scalar field action, and the longitudinal fluctuations are given by a Yang-Mills theory. The induced equation of state has pressure $`p0`$ <sup>4</sup><sup>4</sup>4Note that the above result is still valid when brane fluctuations and fields are large ..
To find the equation of state for the winding modes, we use equation (3.5) to get
$$\stackrel{~}{p}=w_p\rho \mathrm{with}w_p=\frac{p}{d}$$
(3.12)
where $`d`$ is the number of spatial dimensions (9 in our case), and where $`\stackrel{~}{p}`$ and $`\rho `$ stand for the pressure and energy density, respectively.
Fluctuations of the branes and brane matter are given by free scalar and gauge fields on the branes. These may be viewed as particles in the transverse directions extended in brane directions. Therefore, the equation of state is simply that of ordinary matter,
$$\stackrel{~}{p}=w\rho \mathrm{with}\mathrm{\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}0}w1.$$
(3.13)
From the action (3.5) we see that the energy in the winding modes will be
$$E_p(a)T_pa(\eta )^p,$$
(3.14)
where the constant of proportionality is dependent on the number of branes.
The equations of motion for the background are given by
$`d\dot{\lambda }^2+\dot{\phi }^2`$ $`=`$ $`e^\phi E`$ (3.15)
$`\ddot{\lambda }\dot{\phi }\dot{\lambda }`$ $`=`$ $`{\displaystyle \frac{1}{2}}e^\phi P`$ (3.16)
$`\ddot{\phi }d\dot{\lambda }^2`$ $`=`$ $`{\displaystyle \frac{1}{2}}e^\phi E,`$ (3.17)
where $`E`$ and $`P`$ denote the total energy and pressure, respectively,
$$\lambda (t)=log(a(t)),$$
(3.18)
and $`\phi `$ is a shifted dilaton field which absorbs the space volume factor
$$\phi =\mathrm{\hspace{0.17em}2}\varphi d\lambda .$$
(3.19)
The matter sources $`E`$ and $`P`$ are made up of all the components of the brane gas:
$`E`$ $`=`$ $`{\displaystyle \underset{p}{}}E_p^w+E^{nw}`$
$`P`$ $`=`$ $`{\displaystyle \underset{p}{}}w_pE_p^w+wE^{nw},`$ (3.20)
where the superscripts $`w`$ and $`nw`$ stand for the winding modes and the non-winding modes, respectively. The contributions of the non-winding modes of all branes have been combined into one term. The constants $`w_p`$ and $`w`$ are given by (3.12) and (3.13). Each $`E_p^w`$ is the sum of the energies of all of the brane windings with fixed $`p`$.
We may now draw the comparison between the ABE mechanism and . First of all we see that both t-duality and limiting Hagedorn temperature are still manifest once we include the $`p`$-branes . Therefore, there is no physical singularity as $`R0`$. What about the de-compactification mechanism described in section (3.3)? Recall that our initial conditions are in a hot, dense regime near the self dual point $`R=1`$. All the modes (winding, oscillatory and momentum) of all the $`p`$-branes will be excited. By symmetry, we assume that there are equal numbers of winding and anti-winding modes in the system and hence the total winding numbers cancel as in .
Now assume that the universe begins to expand in all directions. The total energy in the winding modes increases with $`\lambda `$ as (3.14), so the largest $`p`$-branes contribute the most. The classical counting argument discussed in is easily generalized to our model. When winding modes meet anti-winding modes, the branes unwind (recall Fig. (3.4)) and allow a certain number of dimensions to grow large.
Consider the probability that the world-volumes of two $`p`$-branes in spacetime will intersect. The winding modes of $`p`$-branes are likely to interact in at most $`2p+1`$ spatial dimensions. <sup>5</sup><sup>5</sup>5To see this, consider the example of two particles (0-branes) moving through a space of dimension $`d`$. These particles will definitely interact (assuming the space is periodic) if $`d=1`$, whereas they probably will not find each other in a space with $`d>1`$.
Since we are in $`d=9`$ spatial dimensions the $`p=8,\mathrm{\hspace{0.17em}6},\mathrm{\hspace{0.17em}5},\mathrm{\hspace{0.17em}4}`$ branes will interact and hence unwind very quickly. For $`p<4`$ a hierarchy of dimensions will be allowed to grow large. Since the energy contained in the winding modes of $`2`$ branes is larger than that of strings (see (3.14)) the $`2`$ branes will have an important effect first. The membranes will allow a $`T^5`$ subspace to grow large. Within this $`5`$ dimensional space the $`1`$-branes will allow a $`T^3`$ subspace to become large. We therefore reach the conclusion that the inclusion of D-branes into the spectrum of fundamental objects in the theory will cause a hierarchy of subspaces to become large while maintaining the results of the BV scenario, explaining the origin of our $`3+1`$-dimensional universe.
Let us summarize the evolution of the ABE universe. The universe starts out in an initial state close to the self-dual point ($`R=1`$), a 9-dimensional toroidal space, hot, dense and filled with particles, strings and $`p`$-brane gases. The universe then starts to expand according to the background equations of motion (3.15 \- 3.17). Branes with the largest value of $`p`$ will have an effect first, and space can only expand further if the winding modes annihilate. The $`8,\mathrm{\hspace{0.17em}6},\mathrm{\hspace{0.17em}5}`$ and $`4`$-branes winding modes annihilate quickly, followed by the $`2`$-branes which allow only $`5`$ spatial dimensions to become large. In this $`T^5`$ the strings allow a $`3`$-dimensional subspace to become large. Hence, it is reasonable to hypothesize the existence of a $`5+1`$-dimensional effective theory at some point in the early history of the universe. In particular, one is tempted to draw a relation between this $`5`$-dimensional picture and the scenario of large extra dimensions proposed in .
There are several problems with the toy model analyzed above. Most of these have already been mentioned by the authors. First, the strings and branes are treated classically. Quantum effects will cause the strings to take on a small but finite thickness , although in our case we are restricted to energy densities lower than the typical string density, and hence the effective width of the strings is of string scale . This presumably will also apply to the branes, although there is no current, consistent quantization scheme developed for branes.
In this scenario there is a brane problem . This is a new problem for cosmological theories with stable branes analogous to the domain wall problem in cosmological scenarios based on quantum field theories with stable domain walls. However, we have found background solutions in our models which approach a point of loitering . Loitering occurs if at some point in the evolution of the universe the size of the Hubble radius extends larger than the physical radius. Such a phase in the background cosmological evolution will naturally solve the brane problem.
The toroidal topology of the compactified manifold was chosen for simplicity. It is important from the point of view of string theory to consider how things would change if this manifold was a Calabi-Yau space. Calabi-Yau three-folds do not admit one cycles for strings to wrap around, although they are necessary if the four-dimensional low energy effective theory is to have $`N=1`$ supersymmetry. Note that in cosmology we do not necessarily expect $`N=1`$ supersymmetry. In particular, maximal supersymmetry is consistent with the toriodal background used.
Also, it was argued in that M-theory should not be formulated in a spacetime of definite dimension or signature. In other words, we must ultimately be able to explain why there is only one time dimension.
Although there is no horizon problem present in this scenario since the universe starts out near the string length and hence there are no causally disconnected regions of space, other problems solved by inflation such as the flatness and structure formation problems are still present. Other less significant concerns are stressed in . This scenario provides a new method for studying string cosmology which is similar to the SBB model and utilizes $`p`$-branes in a very different way from scenarios involving large extra dimensions.
## 4 Pre-Big-Bang
The next attempt to marry cosmology with string theory we will review was proposed in the early 1990s by Veneziano and Gasperini -.
### 4.1 Introduction
The Pre-Big-Bang (PBB) model <sup>6</sup><sup>6</sup>6For an updated collection of papers on this model see http://www.to.infn.it/$``$gasperin. is based on the low energy effective action of string theory, which in $`d`$ spatial dimensions is given by
$$S=\frac{1}{2\lambda _s^{d1}}d^{d+1}x\sqrt{g}e^\phi \left[R+(_\mu \phi )^2+\mathrm{}\right],$$
(4.1)
where $`\phi `$ is the dilaton and $`\lambda _s`$ is the string length scale. The qualitative differences between the PBB model, and the SBB model based on the Einstein-Hilbert action,
$$S=\frac{1}{2\lambda _p^{d1}}d^{d+1}x\sqrt{g}R,$$
(4.2)
are most easily visualized by plotting the history of the curvature of the universe (see Fig. (4.1)) according to each theory. In the SBB scenario the curvature increases as we go back in time, eventually reaching an infinite value at the Big-Bang singularity. In standard inflationary models the curvature reaches some fixed value as $`t`$ decreases at which point the universe enters a de Sitter phase. It has been shown however that such an inflationary phase cannot last forever, for reasons of geodesic completeness, and that the initial singularity problem still remains . The cosmology generated by (4.1) differs drastically from the standard scenarios. The action (4.1) without the โ$`\mathrm{}`$โ terms does not realize the PBB scenario, as we will discuss below. In the PBB model, as one travels back in time the curvature increases as in the previously mentioned models, but in the PBB a maximum curvature is reached at which point the curvature and temperature actually begin to decrease. Although we will examine the details of how this occurs below, a few simple considerations make us feel more comfortable with this picture.
For one, string theory predicts a natural cut-off length scale,
$$\lambda _s=\sqrt{\frac{\mathrm{}}{T}}10l_{pl}10^{32}\text{cm},$$
(4.3)
where $`T`$ is the string tension and $`l_{pl}`$ is the Planck length. So it is natural from the point of view of strings to expect a maximum possible curvature. Logically, as we travel back in time there are only two possibilities if we want to avoid the initial singularity. Either the curvature starts to grow again before the de Sitter phase, in which case we are still left with a singularity shifted earlier in time, or the curvature begins to decrease again, which is what happens in the PBB scenario (Fig. (4.1)c). This behavior is a consequence of scale-factor duality.
> Figure 4.1: Curvature plotted versus time for, (a) the SBB model, (b) the standard inflationary model and (c) the PBB scenario.
### 4.2 More on Duality
To demonstrate the enhanced symmetries present in the PBB model we will examine the consequences of scale-factor duality. The Einstein-Hilbert action (4.2) is invariant under time reversal. Hence, for every solution $`a(t)`$ there exists a solution $`a(t)`$. Or in terms of the Hubble parameter $`H(t)=\dot{a}(t)/a(t)`$, for every solution $`H(t)`$ there exists a solution $`H(t)`$. Thus, if there is a solution representing a universe with decelerated expansion and decreasing curvature ($`H>0`$, $`\dot{H}<0`$) there is a โmirrorโ solution corresponding to a contracting universe ($`H(t)`$, $`H<0`$).
The action of string theory (4.1) is not only invariant under time reversal, but also under inversion of the scale factor $`a(t)`$, (with an appropriate transformation of the dilaton). For every cosmological solution $`a(t)`$ there is a solution $`\stackrel{~}{a}=1/a(t)`$, provided the dilaton is rescaled, $`\phi \stackrel{~}{\phi }=\phi 2d\mathrm{ln}a`$. Hence, time reversal symmetry together with scale-factor duality imply that every cosmological solution has four branches, Fig. (4.2). For the standard scenario of decelerated expansion and decreasing curvature ($`H(t)>0`$, $`\dot{H}(t)<0`$ ) there is a dual partner solution describing a universe with accelerated expansion parameter $`\stackrel{~}{H}(t)`$ and growing curvature $`\dot{\stackrel{~}{H}}(t)`$.
> Figure 4.2: The four branches of a string cosmological solution resulting from scale-factor duality and time reversal.
We will now show how one can create a universe from the string theory perturbative vacuum, that today looks like the standard cosmology. This problem is analogous to finding a smooth way to connect the Pre-Big-Bang phase with a Post-Big-Bang phase, or how to successfully connect the upper-left side of Fig. (4.2) to the upper-right side. In general, the two branches are separated by a future/past singularity and it appears that in order to smoothly connect the branches of growing and decreasing curvature one requires the presence of higher order loop and/or derivative corrections to the effective action (4.1). This cancer of the PBB model is know as the Graceful Exit Problem (GEP) and is the subject of many research papers (see for a collection of references).
One example of how the GEP can be solved is given in . In this work we consider a theory obtained by adding to the usual string frame dilaton gravity action specially constructed higher derivative terms motivated by the limited curvature construction of . The action is (4.1) with the โ$`\mathrm{}`$โ term being replaced by the constructed higher derivative terms. In this scenario all solutions of the resulting theory of gravity are nonsingular and for initial conditions inspired by the PBB scenario solutions exist which smoothly connect a โsuperinflationaryโ phase with $`\dot{H}>0`$, to an expanding FRW phase with $`\dot{H}<0`$, solving the GEP in a natural way.
### 4.3 PBB-Cosmology
Here we examine cosmological solutions of the PBB model. By adding matter in the form of a perfect fluid to the effective action (4.1) (without the โ$`\mathrm{}`$โ terms) and taking a Friedmann-Robertson-Walker background with $`d=3`$, we vary the action to get the equations of motion for string cosmology,
$`\dot{\phi }^26H\dot{\phi }+6H^2`$ $`=`$ $`e^\phi \rho ,`$ (4.4)
$`\dot{H}H\dot{\phi }+3H^2`$ $`=`$ $`{\displaystyle \frac{1}{2}}e^\phi p,`$
$`2\ddot{\phi }+6H\dot{\phi }\dot{\phi }^26\dot{H}12H^2`$ $`=`$ $`0.`$
As an example, for $`p=\rho /3`$ the equations with constant dilaton are exactly solved by
$$at^{1/2},\rho a^4,\phi =\text{const.},$$
(4.5)
which is the standard scenario for the radiation dominated epoch, having decreasing curvature and decelerated expansion:
$$\dot{a}>0,\ddot{a}<0,\dot{H}<0.$$
(4.6)
But there is also a solution obtained from the above via time translation and scale-factor duality,
$$tt,a(t)^{1/2},\phi 3\mathrm{ln}(t),\rho =3pa^2.$$
(4.7)
This solution corresponds to an accelerated, inflationary expansion, with growing dilaton and growing curvature:
$$\dot{a}>0,\ddot{a}>0,\dot{H}>0.$$
(4.8)
Solutions with such behavior are called โsuperinflationaryโ and are located in the upper left quadrant of Fig. (4.2).
Let us briefly review the history of the universe as predicted by the PBB scenario. Recall, that in the SBB model the universe starts out in a hot, dense and highly curved regime. In contrast, the PBB universe has its origins in the simplest possible state we can think of, namely the string perturbative vacuum. Here the universe consists only of a sea of dilaton and gravitational waves. It is empty, cold and flat, which means that we can still trust calculations done with the classical, low-energy effective action of string theory.
In , the authors showed that in a generic case of the PBB scenario, the universe at the onset of inflation must already be extremely large and homogeneous. In order for inflation to solve flatness problems the initial size of a homogeneous part of the universe before PBB inflation must be greater than $`10^{19}l_s`$. In response, it was proposed in that the initial state of the PBB model is a generic perturbative solution of the tree-level, low-energy effective action. Presumably, quantum fluctuations lead to the formation of many black holes (Fig. (4.3)) in the gravi-dilaton sector (in the Einstein frame). Each such singular space-like hypersurface of gravitational collapse becomes a superinflationary phase in the string frame . After the period of dilaton-driven inflation the universe evolves in accordance with the SBB model.
> Figure 4.3: A $`2+1`$ dimensional slice of the string perturbative vacuum giving rise to black hole formation in the Einstein frame.
To conclude let us mention a few benefits of the PBB scenario. For one, there is no need to invent inflation, or fine tune a potential for the inflaton. This model provides a โstringyโ realization of inflation which sets in naturally and is dilaton driven. Pair creation (quantum instabilities) provides a mechanism to heat up an initially cold universe in order to produce a hot big-bang with homogeneity, isotropy and flatness. This scenario also has observable consequences.
Problems with this scenario include the graceful exit problem, mentioned above. This is the problem of smoothly connecting the phases of growing and decreasing curvature, a process that is not well understood and requires further investigation. Most cosmological models require a potential for the dilaton to be introduced by hand in order to freeze the dilaton at late times. In general it is believed that the dilaton should be massive today, otherwise we would notice its effects on physical gauge couplings.
Inclusion of a non-vanishing $`B_{\mu \nu }`$ into the action (4.1) greatly reduces the initial conditions which give rise to inflation . Also the initial collapsing region must be sufficiently large and weakly coupled. Lastly, the dimensionality problem is still present in this model.
## 5 Cosmology and Heterotic M-Theory
In this section we will focus on the work of Lukas, Ovrut and Waldram (LOW) in 1998, which is based on the heterotic M-theory of Hoลava and Witten . Their motivation was to see if it is possible to construct a realistic, cosmological model starting from the most fundamental theory we know.
### 5.1 Hoลava-Witten Theory
In 1996, Hoลava and Witten showed that eleven-dimensional M-theory compactified on an $`S^1/Z_2`$ orbifold with a set of $`E_8`$ gauge supermultiplets on each ten-dimensional orbifold fixed plane can be identified with strongly coupled $`E_8\times E_8`$ heterotic string theory. The basic setup is that of Fig. (5.1), where the orbifold is in the $`x^{11}`$ direction and $`x^{11}[\pi \rho ,\pi \rho ]`$ with the endpoints being identified. The orbifolding with $`Z_2`$ leads to the symmetry $`x^{11}x^{11}`$. It has been shown that this M-theory limit can be consistently compactified on a deformed Calabi-Yau three-fold resulting in an $`N=1`$ supersymmetric theory in four dimensions (see fig.(5.2)). In order to match (at tree level) the gravitational and grand-unified gauge couplings one finds the requirement $`R_{orb}>R_{CY}`$, where $`R_{orb}`$ is the radius of the orbifold and $`R_{CY}10^{16}\text{GeV}`$ is the radius of the Calabi-Yau space. This picture leads to the conclusion that the universe may have gone through a phase in which it was effectively five-dimensional, and therefore provides us with a previously unexplored regime in which to study the early universe.
> Figure 5.1: The Hoลava-Witten scenario. One of the eleven-dimensions has been compactified onto the orbifold $`S^1/Z_2`$. The manifold is $`=IR^{10}\times S^1/Z_2`$.
Here we construct the five-dimensional effective theory via reduction of Hoลava-Witten on a Calabi-Yau three-fold, and then show how this can lead to a four-dimensional toy model for a Friedmann-Robertson-Walker (FRW) universe.
We start with an eleven-dimensional action with bosonic contribution
$$S=S_{SUGRA}+S_{YM},$$
(5.1)
where $`S_{SUGRA}`$ is the action of eleven-dimensional supergravity
$`S_{SUGRA}`$ $`=`$ $`{\displaystyle \frac{1}{2\kappa ^2}}{\displaystyle _{M^{11}}}\sqrt{g}[R+{\displaystyle \frac{1}{24}}G_{IJKL}G^{IJKL}`$ (5.2)
$`+{\displaystyle \frac{\sqrt{2}}{1728}}ฯต^{I_1\mathrm{}I_{11}}C_{I_1I_2I_3}G_{I_4\mathrm{}I_7}G_{I_8\mathrm{}I_{11}}],`$
and $`S_{YM}`$ are two $`E_8`$ Yang-Mills theories on the ten-dimensional orbifold planes
$`S_{YM}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi \kappa ^2}}\left({\displaystyle \frac{\kappa }{4\pi }}\right)^{2/3}{\displaystyle _{M_{10}^{(1)}}}\sqrt{g}\left\{tr(F^{(1)})^2{\displaystyle \frac{1}{2}}trR^2\right\}`$ (5.3)
$`{\displaystyle \frac{1}{8\pi \kappa ^2}}\left({\displaystyle \frac{\kappa }{4\pi }}\right)^{2/3}{\displaystyle _{M_{10}^{(2)}}}\sqrt{g}\left\{tr(F^{(2)})^2{\displaystyle \frac{1}{2}}trR^2\right\}.`$
The values of $`I,J,K,\mathrm{}=0,\mathrm{},9,11`$ parametrize the full eleven-dimensional space $`M_{11}`$, while $`\overline{I},\overline{J},\overline{K},\mathrm{}=0,\mathrm{},9`$ are used for the ten-dimensional hyperplanes, $`M_{10}^{(i)}`$, $`i=1,2`$, orthogonal to the orbifold. The $`F_{\overline{I}\overline{J}}^{(i)}`$ are the two $`E_8`$ gauge field strengths and $`C_{IJK}`$ is the 3-form with field strength given by $`G_{IJKL}=24_{[I}C_{JKL]}`$. In order for this theory to be supersymmetric and anomaly free the Bianchi identity for $`G`$ must pick up the following correction,
$$(dG)_{11\overline{I}\overline{J}\overline{K}\overline{L}}=\frac{1}{2\sqrt{2}\pi }\left(\frac{\kappa }{4\pi }\right)^{2/3}\left\{J^{(1)}\delta (x^{11})+J^{(2)}\delta (x^{11}\pi \rho )\right\}_{\overline{I}\overline{J}\overline{K}\overline{L}}$$
(5.4)
where the sources are
$$J^{(i)}=trF^{(i)}F^{(i)}\frac{1}{2}trRR.$$
(5.5)
Now, we search for solutions to the above theory which preserve four of the thirty-two supercharges and, when compactified, lead to four dimensional, $`N=1`$ supergravities. To begin, consider the manifold $`=IR^4\times X\times S^1/Z_2`$, where $`IR^4`$ is four-dimensional Minkowski space and $`X`$ is a Calabi-Yau three-fold. Upon compactification onto $`X`$, we are left with a five-dimensional effective spacetime consisting of two copies of $`IR^4`$, one at each of the orbifold fixed points, and the orbifold itself (see fig. (5.2)). On each of the $`IR^4`$ planes there is a gauge group $`H^{(i)}`$, $`i=1,2`$, and $`N=1`$.
> Figure 5.2: The LOW scenario. The manifold is given by $`=IR^4\times X\times S^1/Z_2`$ where the Hoลava-Witten theory is compactified on a smooth Calabi-Yau three fold $`X`$. Compactification results in a five-dimensional effective theory.
In the next section we construct the five-dimensional effective theory.
### 5.2 Five-Dimensional Effective Theory
As we have discussed, according to the model presented above, there is an epoch when the universe appears to be five dimensional. Hence, it is only natural to try to find the action for this five-dimensional effective theory. Let us identify the fields in the five-dimensional bulk. First, there is the gravity multiplet $`(g_{\alpha \beta },๐_\alpha ,\psi _\alpha ^i)`$, where $`g_{\alpha \beta }`$ is the graviton, $`๐_\alpha `$ is a five-dimensional vector field, and the $`\psi _\alpha ^i`$ are the gravitini. The indices $`\alpha ,\beta =0,\mathrm{},3,11`$ and $`i=1,2`$. There is also the universal hypermultiplet $`q(V,\sigma ,\xi ,\overline{\xi },\zeta ^i)`$. Here $`V`$ is a modulus field associated with the volume of the Calabi-Yau space, $`\xi `$ is a complex scalar zero mode, $`\sigma `$ is a scalar resulting from the dualization of the three-form $`C_{\alpha \beta \gamma }`$, and the $`\zeta ^i`$ are the hypermultiplet fermions.
It is now possible, using the action (5.1) to construct the five-dimensional effective action of Hoลava-Witten theory,
$$S_5=S_{grav}+S_{hyper}+S_{bound},$$
(5.6)
where,
$$S_{grav}=\frac{v}{2\kappa ^2}_{M_5}\sqrt{g}\left[R+\frac{3}{2}(_{\alpha \beta })^2+\frac{1}{\sqrt{2}}ฯต^{\alpha \beta \gamma \delta ฯต}๐_\alpha _{\beta \gamma }_{\delta ฯต}\right],$$
(5.7)
$$S_{hyper}=\frac{v}{2\kappa ^2}_{M_5}\sqrt{g}\left[4h_{\mu \nu }_\alpha q^\mu ^\alpha q^\nu +\frac{\alpha ^2}{3}V^2\right],$$
(5.8)
$`S_{bound}`$ $`=`$ $`{\displaystyle \frac{v}{2\kappa ^2}}\left[2\sqrt{2}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle _{M_4^{(i)}}}\sqrt{g}\alpha V^1\right]`$ (5.9)
$`{\displaystyle \frac{v}{8\pi \kappa ^2}}\left({\displaystyle \frac{\kappa }{4\pi }}\right)^{2/3}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle _{M_4^{(i)}}}\sqrt{g}Vtr(F_{\mu \nu }^{(i)})^2.`$
In the above, $`v`$ is a constant that relates the five-dimensional Newton constant, $`\kappa _5`$, with the eleven-dimensional Newton constant, $`\kappa `$, via $`\kappa _5^2=\kappa ^2/v`$. The metric $`h_{\mu \nu }`$ is the flat space metric and $`\alpha `$ is a constant. Higher-derivative terms have been dropped and this action provides us with a minimal $`N=1`$ supergravity theory in the five-dimensional bulk.
This theory admits a three-brane domain wall solution with a world-volume lying in the four uncompactified dimensions . In fact, a pair of domain walls is the vacuum solution of the five-dimensional theory which provides us with a background for reduction to a $`d=4`$, $`N=1`$ effective theory. This solution will be the topic of the next section.
### 5.3 Three-Brane Solution
In order to find a pair of three-branes solution we should start with an ansatz for the five-dimensional metric of the form
$`ds_5^2`$ $`=`$ $`a(y)^2dx^\mu dx^\nu \eta _{\mu \nu }+b(y)^2dy^2`$ (5.10)
$`V`$ $`=`$ $`V(y),`$
where $`y=x^{11}`$. By using the equations of motion derived from the action (5.6) we find
$`a`$ $`=`$ $`a_0H^{1/2}`$ (5.11)
$`b`$ $`=`$ $`b_0H^2`$
$`V`$ $`=`$ $`b_0H^3,`$
where $`H\frac{\sqrt{2}}{3}\alpha _0|y|+c_0`$, and $`a_0,b_0`$ and $`c_0`$ are all constants. Using the equations of motion derived by varying the action with respect to $`g_{\mu \nu }`$ of (5.10), we arrive at a differential equation which leads to
$$_y^2H=\frac{2\sqrt{2}}{3}\alpha _0\left(\delta (y)\delta (y\pi \rho )\right).$$
(5.12)
A detailed derivation of this equation is discussed in . Clearly, (5.12) represents two parallel three-branes located at the orbifold planes, as in Fig.(5.2). This solves the five-dimensional theory exactly and preserves half of the supersymmetries, with low-energy gauge and matter fields carried on the branes. This prompts us to find realistic cosmological models from the above scenario where the universe lives on the world-volume of a three-brane.
### 5.4 Cosmological Domain-Wall Solution
In order to construct a dynamical, cosmological solution, the solutions in (5.11) are made to be functions of time $`\tau `$, as well as the eleventh dimension $`y`$,
$`ds_5^2`$ $`=`$ $`N(\tau ,y)d\tau ^2+a(\tau ,y)^2dx^mdx^n\eta _{mn}+b(\tau ,y)^2dy^2`$ (5.13)
$`V`$ $`=`$ $`V(\tau ,y).`$
Here we have introduced a lapse function $`N(\tau ,y)`$. Because this ansatz leads to a very complicated set of non-linear equations we will seek a solution based on the separation of variables. Note, there is no a priori reason to believe that such a solution exists, but we will see that one does. Separating the variables $`\tau `$ and $`y`$,
$`N(\tau ,y)`$ $`=`$ $`n(\tau )a(y)`$ (5.14)
$`a(\tau ,y)`$ $`=`$ $`\alpha (\tau )a(y)`$
$`b(\tau ,y)`$ $`=`$ $`\beta (\tau )b(y)`$
$`V(\tau ,y)`$ $`=`$ $`\gamma (\tau )V(y).`$
Since this article is intended only as an elementary review we will not repeat the details involved in solving the above system. For our purposes it suffices to say that the equations take on a particularly simple form when $`\beta =\gamma `$ and with the gauge choice of $`n=const.`$. In this gauge, $`\tau `$ becomes proportional to the comoving time $`t`$, since $`dt=n(\tau )d\tau `$. A solution exists such that
$`\alpha `$ $`=`$ $`A|tt_0|^p`$ (5.15)
$`\beta `$ $`=`$ $`\beta =B|tt_0|^q,`$
where
$`p`$ $`=`$ $`{\displaystyle \frac{3}{11}}(1{\displaystyle \frac{4}{3\sqrt{3}}})`$ (5.16)
$`q`$ $`=`$ $`{\displaystyle \frac{2}{11}}(1\pm 2\sqrt{3}),`$
and $`A,B`$ and $`t_0`$ are arbitrary constants. This is the desired cosmological solution. The $`y`$-dependence is identical to the domain wall solution (5.12) and the scale factors evolve with $`t`$ according to (5.15). The domain wall pair remain rigid, while their sizes and the separation between the walls change. In particular, $`\alpha `$ determines the size of the domain-wall world-volume while $`\beta `$ gives the separation of the two walls. In other words, $`\alpha `$ determines the size of the three-dimensional universe, while $`\beta `$ gives the size of the orbifold. Furthermore, the $`d=4`$ world-volume of the three-brane universe exhibits $`N=1`$ SUSY (of course SUSY is broken in the dyanamical solution) and a particular solution exists for which the domain wall world-volume expands in a FRW-like manner while the orbifold radius contracts.
Although the above model provides an intriguing use of M-theory in an attempt to answer questions about early universe cosmology there are still many problems to be worked out. Foremost, these are vacuum solutions, devoid of matter and radiation. There is no reason to think that, of all the solutions, the one which matches our universe (expanding domain-wall, shrinking orbifold) should be preferred over any other. This problem is typical of many cosmological models, however. The Calabi-Yau (six-dimensional) three-fold is chosen by hand in order to give four noncompact dimensions. Hence, the dimensionality problem mentioned in Section 3 is still present in this model. Stabilization of moduli fields, including the dilaton has recently been addressed in . There are no cosmological constants in the model. There is also no natural mechanism supplied for SUSY breaking on the domain wall, and currently no discussion of inflationary dynamics. For more on heterotic M-theory and cosmology see, -.
## 6 Large Extra Dimensions
This section provides a brief discussion of scenarios involving large extra dimensions, focusing primarily on the models of Randall and Sundrum (RSI and RSII)<sup>7</sup><sup>7</sup>7The distinction between RSI and RSII models will be clarified below. . The RSI model is similar in many respects to that of the Lukas, Ovrut and Waldram scenario discussed in section 5, although its motivation is quite different. In the LOW construction the motivation was to construct a cosmology out of the fundamental theory of everything. In the RSI model the motivation is to construct a cosmology in which the Hierarchy problem of the Standard Model (SM) is solved in a natural way. Some earlier proposals involving large extra dimensions include -. Also see the extensive set of references in .
### 6.1 Motivation and the Hierarchy Problem
There is a hierarchy problem in the Standard Model because we have no way of explaining why the scales of particle physics are so different from those of gravity. Many attempts to solve the hierarchy problem using extra dimensions have been made before, see for example and . If spacetime is fundamentally $`(4+n)`$-dimensional then the physical Planck mass
$$M_{pl}^{(4)}2\times 10^{18}\text{GeV},$$
(6.1)
is actually dependent on the fundamental $`(4+n)`$-dimensional Planck mass $`M_{pl}`$ and on the geometry of the extra dimensions according to
$$M_{pl}^{(4)}{}_{}{}^{2}=M_{pl}^{n+2}V_n,$$
(6.2)
Here $`V_n`$ is the volume of the $`n`$ compact extra dimensions. Because we have not detected any extra dimensions experimentally, the compactification scale $`\mu _c1/V_n^{1/n}`$ would have to be much smaller than the weak scale, and the particles and forces of the SM (except for gravity) must be confined to the four-dimensional world-volume of a three-brane (See Fig. (6.1)).
> Figure 6.1: In the RS model the fields of the Standard Model (with the exception of gravity) are confined to the three-brane world-volume while gravity is allowed to propagate in the bulk.
We see from (6.2) that by taking $`V_n`$ to be large enough it is possible to eliminate the hierarchy between the weak scale $`v`$ and the Planck scale. Unfortunately, in this procedure a new hierarchy has been introduced, namely the one between $`\mu _c`$ and $`v`$. Randall and Sundrum proposed the following: We assume that the particles and forces of the SM with the exception of gravity are confined to a four-dimensional subspace of the $`(4+n)`$-dimensional spacetime. This subspace is identified with the world-volume of a three-brane and an ansatz for the metric is made. Randall and Sundrumโs proposal is that the metric is not factorizable, but the four-dimensional metric is multiplied by a โwarpโ factor that is exponentially dependent upon the radius of the bulk, fifth dimension. The metric ansatz is
$$ds^2=e^{2kr_c\phi }\eta _{\mu \nu }dx^\mu dx^\nu +r_c^2d\phi ^2,$$
(6.3)
where $`k`$ is a scale of order the Planck scale, $`\eta _{\mu \nu }`$ is the four-dimensional Minkowski metric and $`0\phi \pi `$ is the coordinate for the extra dimension. Randall and Sundrum have shown that this metric solves the Einstein equations and represents two three-branes with appropriate cosmological constant terms separated by a fifth dimension. The above scenario, in addition to being able to solve the hierarchy problem (see section 6.2.1), provides distinctive experimental signatures. Coupling of an individual Kaluza-Klein (KK) excitation to matter or to other gravitational modes is set by the weak and not the Planck scale. There are no light KK modes because the excitation scale is of the order a TeV. Hence, it should be possible to detect such excitations at accelerators (such as the LHC). The KK modes are observable as spin $`2`$ excitations that can be reconstructed from their decay products. For experimental signatures of KK modes within large extra dimensions see e.g. .
### 6.2 Randall-Sundrum I
The basic setup for the RSI model is depicted in Fig. (6.2). The angular coordinate $`\phi `$ parameterizes the fifth dimension and ranges from $`\pi `$ to $`\pi `$. The fifth dimension is taken as the orbifold $`S^1/Z_2`$ where there is the identification of $`(x,\phi )`$ with $`(x,\phi )`$. The orbifold fixed points are at $`\phi =0,\pi `$ and correspond with the locations of the three-brane boundaries of the five-dimensional spacetime. Note the similarities of this model with the LOW model of Section 5. One difference is that we are now considering nonzero vacuum energy densities on both the visible and the hidden brane and in the bulk.
> Figure 6.2: The Randall-Sundrum scenario. The fifth dimension is compactified onto the orbifold $`S^1/Z_2`$.
The action describing the scenario is
$$S=S_{grav}+S_{vis}+S_{hid}$$
(6.4)
where
$`S_{grav}`$ $`=`$ $`{\displaystyle d^4x_\pi ^\pi ๐\phi \sqrt{G}\left(\mathrm{\Lambda }+2M^3R\right)}`$
$`S_{vis}`$ $`=`$ $`{\displaystyle d^4x\sqrt{g_{vis}}\left(_{vis}+V_{vis}\right)}`$
$`S_{hid}`$ $`=`$ $`{\displaystyle d^4x\sqrt{g_{hid}}\left(_{hid}+V_{hid}\right)}.`$ (6.5)
Here, $`M`$ is the Planck mass, $`R`$ is the Ricci scalar, $`g_{vis}`$ and $`g_{hid}`$ are the four-dimensional metrics on the visible and hidden sectors respectively and $`V_{vis}`$, $`\mathrm{\Lambda }`$ and $`V_{hid}`$ are the cosmological constant terms in the visible, bulk and hidden sectors. The specific form for the three-brane Lagrangians is not relevant for finding the classical five-dimensional, ground state metric. The five-dimensional Einstein equations for the above action are
$`\sqrt{G}\left(R_{MN}{\displaystyle \frac{1}{2}}G_{MN}R\right)`$ $`=`$ $`{\displaystyle \frac{1}{4M^3}}[\mathrm{\Lambda }\sqrt{G}G_{MN}`$ (6.6)
$`+V_{vis}\sqrt{g_{vis}}g_{\mu \nu }^{vis}\delta _M^\mu \delta _N^\nu \delta (\phi \pi )`$
$`+V_{hid}\sqrt{g_{hid}}g_{\mu \nu }^{hid}\delta _M^\mu \delta _N^\nu \delta (\phi ).`$
We now assume that a solution exists which has four-dimensional Poincarรฉ invariance in the $`x^\mu `$ directions. A five-dimensional ansatz which obeys the above requirements is
$$ds^2=e^{2\sigma (\phi )}\eta _{\mu \nu }dx^\mu dx^\nu +r_c^2d\phi ^2.$$
(6.7)
Substituting this ansatz into (6.6) reduces the Einstein equations to
$`{\displaystyle \frac{6\sigma ^2}{r_c^2}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Lambda }}{4M^3}},`$ (6.8)
$`{\displaystyle \frac{3\sigma ^{\prime \prime }}{r_c^2}}`$ $`=`$ $`{\displaystyle \frac{V_{hid}}{4M^3r_c}}\delta (\phi )+{\displaystyle \frac{V_{vis}}{4M^3r_c}}\delta (\phi \pi ).`$ (6.9)
Solving (6.8) consistently with orbifold symmetry $`\phi \phi `$, we find
$$\sigma =r_c|\phi |\sqrt{\frac{\mathrm{\Lambda }}{24M^3}},$$
(6.10)
which makes sense if $`\mathrm{\Lambda }<0`$. With this choice, the spacetime in the bulk of the theory is a slice of an $`AdS_5`$ manifold. Also, to solve (6.9) we should take
$`V_{hid}=V_{vis}`$ $`=`$ $`24M^3k`$
$`\mathrm{\Lambda }`$ $`=`$ $`24M^3k^2.`$ (6.11)
Note that the boundary and bulk cosmological terms are dependent upon the single scale factor $`k`$, and that the relations between them are required in order to get four-dimensional Poincarรฉ invariance.
Further connections with the LOW scenario of Section 5 are now visible. The exact same relations given in (6.2) arise in the five-dimensional Hoลava-Witten effective theory if one identifies the expectation values of the background three-form field as cosmological terms .
We want the bulk curvature to be small compared to the higher dimensional Planck scale in order to trust the solution and thus, we assume $`k<M`$. The bulk metric solution is therefore,
$$ds^2=e^{2kr_c|(\phi )|}\eta _{\mu \nu }dx^\mu dx^\nu +r_c^2d\phi ^2.$$
(6.12)
Since $`r_c`$ is small but still larger than $`1/k`$, the fifth dimension cannot be experimentally observed in present or future gravity experiments. This prompts us to search for a four-dimensional effective theory.
#### 6.2.1 Four-Dimensional Effective Theory
In our four-dimensional effective description we wish to find the parameters of this low-energy theory (e.g. $`M_{pl}^{(4)}`$ and mass parameters of the four-dimensional fields ) in terms of the five-dimensional, fundamental scales, $`M`$, $`k`$ and $`r_c`$. In order to find the four-dimensional theory one identifies massless gravitational fluctuations about the classical solution 6.12 which correspond to the gravitational fields for the effective theory. These are the zero modes of the classical solution. The metric of the four-dimensional effective theory is of the form
$$\overline{g}_{\mu \nu }(x)\eta _{\mu \nu }+\overline{h}_{\mu \nu }(x),$$
(6.13)
which is locally Minkowski. Here, $`\overline{h}_{\mu \nu }(x)`$ represents the tensor fluctuations about Minkowski space and gives the physical graviton of the four-dimensional effective theory. By substituting the metric (6.13) for $`\eta _{\mu \nu }`$ in (6.12) and then using the result in the action (6.2) the curvature term becomes
$$S_{eff}d^4x_\pi ^\pi ๐\phi \mathrm{\hspace{0.17em}2}M^3r_ce^{2kr_c|\phi |}\sqrt{\overline{g}}\overline{R},$$
(6.14)
where $`\overline{R}`$ is the four-dimensional Ricci scalar made out of $`\overline{g}_{\mu \nu }(x)`$. We focus on the curvature term so that we may derive the scale of the gravitational interactions. The effective fields depend only on $`x`$, and hence it is possible to perform the integration over $`\phi `$ explicitly, obtaining the four-dimensional effective theory . Using the result one may derive an expression for the four-dimensional Planck mass in terms of the fundamental, five-dimensional Planck mass
$$M_{pl}^{(4)}{}_{}{}^{2}=M^3r_c_\pi ^\pi ๐\phi e^{2kr_c|\phi |}=\frac{M^3}{k}[1e^{2kr_c\pi }].$$
(6.15)
Notice that $`M_{pl}^{(4)}`$ depends only weakly on $`r_c`$ in the large $`kr_c`$ limit.
From the fact that $`g_{\mu \nu }^{vis}(x^\mu )G_{\mu \nu }(x^\mu ,\phi =\pi )`$ and $`g_{\mu \nu }^{hid}(x^\mu )G_{\mu \nu }(x^\mu ,\phi =0)`$ we find,
$$\overline{g}_{\mu \nu }=g_{\mu \nu }^{hid},$$
(6.16)
but
$$\overline{g}_{\mu \nu }=g_{\mu \nu }^{vis}e^{2kr_c\pi }.$$
(6.17)
It is now possible to find the matter field Lagrangian of the theory. With proper normalization of the fields one can determine physical masses. Let us consider the example of a fundamental Higgs field. The action is
$$S_{vis}d^4x\sqrt{g_{vis}}\left(g_{vis}^{\mu \nu }D_\mu H^{}D_\nu H\lambda \left(|H|^2v_0^2\right)^2\right),$$
(6.18)
which contains only one mass parameter $`v_0`$. Using (6.17) the action becomes
$$S_{eff}d^4x\sqrt{\overline{g}}e^{4kr_c\pi }\left(\overline{g}^{\mu \nu }e^{2kr_c\pi }D_\mu H^{}D_\nu H\lambda \left(|H|^2v_0^2\right)^2\right),$$
(6.19)
and after wavefunction renormalization, $`He^{kr_c\pi }H`$, we have
$$S_{eff}d^4x\sqrt{\overline{g}}\left(\overline{g}^{\mu \nu }D_\mu H^{}D_\nu H\lambda \left(|H|^2e^{2kr_c\pi }v_0^2\right)^2\right).$$
(6.20)
This result is completely general. The physical mass scales are set by a symmetry-breaking scale,
$$ve^{kr_c\pi }v_0,$$
(6.21)
and hence any mass parameter $`m_0`$ on the visible three-brane is related to the fundamental, higher-dimensional mass via
$$me^{kr_c\pi }m_0.$$
(6.22)
Note that if $`e^{kr_c\pi }10^{15}`$, TeV scale physical masses are produced from fundamental mass parameters near the Planck scale, $`10^{19}\text{GeV}`$. Therefore, there are no large hierarchies if $`kr_c50`$.
### 6.3 Randall-Sundrum II
In the RSI scenario described in the last section our universe was identified with the negative tension brane while the brane in the hidden sector had positive tension (Eq. (6.2)). In this model it was shown that the hierarchy problem may be solved. Unfortunately, there are several problems with the idea that the universe we live in can be a negative tension brane. For one, the energy density of matter on such a brane would be negative and gravity repulsive . Life is more comfortable on a positive tension brane since the D-branes which arise as fundamental objects in string theories are all positive tension objects and the localization of matter and gauge fields on positive tension branes is well understood within the context of string theory.
For the above reasons Randall and Sundrum suggested a second scenario (RSII) in which our universe is the positive tension brane and the hidden brane has negative tension . In this case the boundary and bulk cosmological constants are related by
$`V_{vis}=V_{hid}`$ $`=`$ $`24M^3k`$
$`\mathrm{\Lambda }`$ $`=`$ $`24M^3k^2,`$ (6.23)
as opposed to the realation in RSI, Eq. (6.2).
As we will see, in this refined scenario it is possible to reproduce Newtonian gravity and other four-dimensional general relativistic predictions at low energy and long distance on the visible brane. Note that in the solution to the hierarchy problem in RSI the wave function for the massless graviton is greatest on the hidden brane, whereas RSII has the graviton bound to the visible brane. To see this, consider the wave equation for small gravitational fluctuations,
$$\left(_\mu ^\mu _i^i+V(z_i)\right)\widehat{h}(x^\mu ,z_i)=0.$$
(6.24)
This has a non-trivial potential term $`V`$ resulting from the curvature, $`\mu `$ runs from $`0`$ to $`3`$ and $`i`$ labels the extra dimensions. It is possible to write $`\widehat{h}`$ as a superposition of modes $`\widehat{h}=e^{ipx}\widehat{\psi }(z)`$ where $`\widehat{\psi }`$ is an eigenmode of the equation
$$\left(_i^i+V(z)\right)\widehat{\psi }(z)=m^2\widehat{\psi }(z),$$
(6.25)
in the extra dimensions and $`p^2=m^2`$. Hence, the higher-dimensional gravitational fluctuations are Kaluza-Klein reduced in terms of four-dimensional KK states with mass $`m^2`$ given by the eigenvalues of (6.25). The zero mode that is also a normalizable state in the spectrum of Eq. (6.25) is the wave function associated with the four-dimensional graviton. This state is a bound state whose wave function falls off rapidly away from the 3-brane. Such behavior corresponds to a 3-brane acting as a positive tension source on the right hand side of Einsteinโs equations.
The procedure of RSII is to decompactify the orbifold of RSI (i.e. consider $`r_c\mathrm{}`$) taking the hidden, negative tension brane off to infinity. In doing this, one obtains an effective four-dimensional theory of gravity where the setup is a single three-brane with positive tension embedded in a five-dimensional bulk spacetime. On this brane one can compute an effective nonrelativistic gravitational potential between two particles of masses $`m_1`$ and $`m_2`$ which is generated by exchange of the zero-mode and continuum Kaluza-Klein mode propagators. The potential behaves as
$$V(r)=G_N\frac{m_1m_2}{r}\left(1+\frac{1}{r^2}k^2\right).$$
(6.26)
Here the leading term is the usual Newtonian potential and is due to the bound state mode. The KK modes generate the $`1/r^3`$ correction term which is heavily suppressed for $`k`$ of order the fundamental Planck scale and $`r`$ of the size tested with gravity. The propagators calculated in are relativistic and hence, going beyond the nonrelativistic approximation one recovers all the proper relativistic corrections with negligible corrections from the continuum modes.
Let us compare the RSI and RSII models. In RSI, the solution to the hierarchy problem requires that we are living on a negative tension brane. The positive tension brane has no such suppression of its masses and is therefore often referred to as the โPlanckโ brane, which is hidden from the visible brane. Serious arguments against this scenario are that the negative tension โTeVโ brane seems physically unacceptable.
In RSII, the visible brane is taken as the positive tension brane while the TeV brane is sent off to infinity. In this model the proper Newtonian gravity is manifest on the visible brane, but the hierarchy problem is not addressed.
Although more successful as a potential physical model of our universe than its predecessor RSI, RSII seems to lack the elegant solution to the hierarchy problem made possible by considering the universe as a negative tension brane. Recent work however suggests that by including quantum effects (analogous to the Casimir effect) it is possible to solve the hierarchy problem on the visible brane having either positive or negative tension . If the Casimir energy is negative and one accepts a degree of fine tuning of the tension on the hidden brane it is possible to obtain a large enough warp factor to explain the hierarchy on the visible brane having either positive or negative tension. Further work on this scenario is needed however including a study of the stability of this model against perturbations.
### 6.4 RS and Brane World Cosmology
The next obvious step is to consider the cosmologies of the RS model discussed above. There has been an extensive amount of work done in these areas and the reader is invited to examine the references at the end of the review related to Randall-Sundrum and โbrane worldโ cosmologies \- for a comprehensive study. Due to the vast number of cosmological models discussed in the literature we will review only the basics and focus on the problems of brane world cosmologies while mentioning potential resolutions and future work, referencing various relevant authors. Much of the discussion in this section closely parallels the excellent review of J. Cline .
We begin by considering the cosmological expansion of 3-brane universes in a 5-dimensional bulk with a cosmological constant as discussed by Binรฉtruy, Deffayet, Ellwanger and Langlois (BDEL) . Note that in an earlier work , BDL considered the solutions to Einsteinโs equations in five dimensions with an $`S_1/Z_2`$ orbifold and matter included on the two branes but with no cosmological constants on the branes or in the bulk. They found that the Hubble expansion rate of the visible brane was related to the energy density of the brane quadratically opposed to the standard Friedmann equation, $`H^2\rho `$. We will show this explicitly below. The altered expansion rate proved to be incompatible with nucleosynthesis constraints.
When the analysis was applied to the RSII scenario one does in fact reproduce the ordinary FRW universe on the positive tension, Planck brane . Note however, that in the RSII scenario on the negative tension brane where the hierarchy problem is solved the Friedmann equation has a critical sign difference.
In the BDEL model the authors consider five-dimensional spacetime metrics of the form
$$ds^2=\stackrel{~}{g}_{AB}dx^Adx^B=g_{\mu \nu }dx^\mu dx^\nu +b^2dy^2$$
(6.27)
where $`y`$ is the coordinate associated with the fifth dimension. The visible universe is taken to be the hypersurface at $`y=0`$. The metric is taken to be
$$ds^2=n^2(\tau ,y)d\tau ^2+a^2(\tau ,y)g_{ij}dx^idx^j+b^2(\tau ,y)dy^2,$$
(6.28)
where $`\gamma _{ij}`$ is a maximally symmetric three-dimensional metric ($`k=1,0,1`$ will parametrize the spatial curvature), and $`n`$ is a lapse function.
The five-dimensional Einstein equations have the usual form
$$\stackrel{~}{G}_{AB}\stackrel{~}{R}_{AB}\frac{1}{2}\stackrel{~}{R}\stackrel{~}{g}_{AB}=\kappa ^2\stackrel{~}{T}_{AB},$$
(6.29)
where $`\kappa `$ is related to the five-dimensional Newtonโs constant $`G_{(5)}`$ and the five-dimensional reduce Planck mass $`M_{(5)}`$ by
$$\kappa ^2=8\pi G_{(5)}=M_{(5)}^3.$$
(6.30)
Using the ansatz (6.28) one finds the non-vanishing components of the Einstein tensor to be
$`\stackrel{~}{G}_{00}`$ $`=`$ $`3\left\{{\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{\dot{b}}{b}}\right){\displaystyle \frac{n^2}{b^2}}\left({\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}{\displaystyle \frac{b^{}}{b}}\right)\right)+k{\displaystyle \frac{n^2}{a^2}}\right\},`$ (6.31)
$`\stackrel{~}{G}_{ij}`$ $`=`$ $`{\displaystyle \frac{a^2}{b^2}}\gamma _{ij}\left\{{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}+2{\displaystyle \frac{n^{}}{n}}\right){\displaystyle \frac{b^{}}{b}}\left({\displaystyle \frac{n^{}}{n}}+2{\displaystyle \frac{a^{}}{a}}\right)+2{\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{n^{\prime \prime }}{n}}\right\}`$ (6.32)
$`+{\displaystyle \frac{a^2}{n^2}}\gamma _{ij}\left\{{\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}+2{\displaystyle \frac{\dot{n}}{n}}\right)2{\displaystyle \frac{\ddot{a}}{a}}+{\displaystyle \frac{\dot{b}}{b}}\left(2{\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{\dot{n}}{n}}\right){\displaystyle \frac{\ddot{b}}{b}}\right\}k\gamma _{ij},`$
$`\stackrel{~}{G}_{05}`$ $`=`$ $`3\left({\displaystyle \frac{n^{}}{n}}{\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{a^{}}{a}}{\displaystyle \frac{\dot{b}}{b}}{\displaystyle \frac{\dot{a}^{}}{a}}\right),`$ (6.33)
$`\stackrel{~}{G}_{55}`$ $`=`$ $`3\left\{{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}+{\displaystyle \frac{n^{}}{n}}\right){\displaystyle \frac{b^2}{n^2}}\left({\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}{\displaystyle \frac{\dot{n}}{n}}\right)+{\displaystyle \frac{\ddot{a}}{a}}\right)k{\displaystyle \frac{b^2}{a^2}}\right\}.`$ (6.34)
Here the prime indicates differentiation with respect to $`y`$ and dot indicates differentiation with respect to $`\tau `$.
The energy-momentum tensor can be described in terms of the fields living on the visible brane world-volume $`T_B^A`$ and the fields living in the bulk space (or on other branes) $`\stackrel{ห}{T}_B^A`$. We have
$$T_B^A=\frac{\delta (y)}{b}diag(\rho ,p,p,p,0),$$
(6.35)
where the energy density $`\rho `$ and pressure $`p`$ are independent of the position in the brane in order to recover a homogeneous cosmology on the brane. The total energy-momentum tensor is then
$$\stackrel{~}{T}_B^A=T_B^A+\stackrel{ห}{T}_B^A.$$
(6.36)
Note that in reality the brane would have some thickness in the fifth dimension determined by the fundamental scale of the underlying theory. However, the presence of the delta function in (6.35) (the โthin-braneโ approximation) should be valid when the energy scales are much smaller than the fundamental scale. In what follows unless otherwise mentioned we will take $`k=0`$ as in .
From the Bianchi identity $`_A\stackrel{~}{G}_B^A=0`$ and the Einstein equations (6.29) an equation of conservation is obtained,
$$\dot{\rho }+3(p+\rho )\frac{\dot{a}_0}{a_0}=0,$$
(6.37)
which matches the usual four-dimensional equation of energy density conservation in standard cosmology. Here $`a_0`$ is the value of $`a`$ on the brane.
To find a solution of Einsteinโs equations (6.29) in the vicinity of the visible brane at $`y=0`$ one must deal with the delta function sources. The details may be found in . From the $`55`$ component equation (6.34) one finds a new Friedmann-like equation
$$\frac{\dot{a}_0^2}{a_0^2}+\frac{\ddot{a}_0}{a_0}=\frac{\kappa _{(5)}^4}{36}\rho (\rho +3p)\frac{\kappa _{(5)}^2\stackrel{ห}{T}_{55}}{3b_0^2}.$$
(6.38)
Immediately, and as mentioned above one sees the unusual quadratic dependence $`H^2\rho ^2`$.
Note that if one allows for a cosmological constant in the bulk $`\mathrm{\Lambda }`$ and extra energy densities on the Planck and TeV branes ($`\rho _P`$ and $`\rho _T`$, respectively), in addition to the respective tensions $`\sigma `$ and $`\sigma `$ one finds a Friedmann-like equation of the form <sup>8</sup><sup>8</sup>8Now we have switched to the notation of but here $`\overline{k}`$ is the $`k`$ introduced in our discussion of RSI, Section (6.2).
$$H^2=\frac{(\sigma +\rho _p)^2}{36M^6}+\frac{\mathrm{\Lambda }}{6M^3}=\frac{(\sigma +e^{2\overline{k}b}\rho _T)^2}{36M^6}+\frac{\mathrm{\Lambda }}{6M^3}.$$
(6.39)
When the tension $`\sigma `$ is fine tuned to cancel the contribution from $`\mathrm{\Lambda }`$ in the limit $`\rho _i=0`$ , it is possible to recover the correct FRW behavior $`H\sqrt{\rho }`$ at leading order in $`\rho `$ . Interestingly, this fine tuning is exactly that required by RS to obtain a static solution. Unfortunately while the cosmology on the Planck brane appeared normal, the energy density on the TeV brane, where the hierarchy problem is solved, is negative which is physically unacceptable as we mentioned above.
Exciting new developments have shown that when the radion is stabilized the previously mentioned unconventional cosmologies in the RS model disappear . By assuming that a 5-dimensional potential $`U(b)`$ is generated by some mechanism (e.g. ) in the 5-dimensional theory, the nonvanishing equations of motion in the bulk (with cosmological constant $`\mathrm{\Lambda }`$) reduce to
$`\stackrel{~}{G}_{00}`$ $`=`$ $`\kappa ^2n^2(\mathrm{\Lambda }+U(b)),`$
$`\stackrel{~}{G}_{ii}`$ $`=`$ $`\kappa ^2a^2(\mathrm{\Lambda }+U(b)),`$
$`\stackrel{~}{G}_{55}`$ $`=`$ $`\kappa ^2b^2(\mathrm{\Lambda }+U(b)+U^{}(b)).`$ (6.40)
Here $`\stackrel{~}{G}_{AB}`$ is given by (6.31)-(6.34) with $`k=0`$. Let us introduce notation $`m_0`$ such that the static RS solution is recovered when $`V_p=V_T=6m_0/\kappa ^2`$ and $`\mathrm{\Lambda }=6m_0^2/\kappa ^2`$. We take the locations of the Planck and TeV branes to be at $`y=0`$ and $`y=1/2`$, respectively. To simplify the solution of (6.4) the radion is assumed to be very heavy and near its minimum $`UM_b^5((bb_0)/b_0)^2`$. Here $`b_0`$ is the stabilized value of $`b`$ and $`M_b`$ is proportional to the radion mass $`m_{rad}`$. Now one may perturb around the RS solution, with cosmological constants $`\delta V_p`$ and $`\delta V_T`$ instead of matter densities. Using the ansatz
$`a(t,y)`$ $`=`$ $`e^{Ht|y|m_0b_0}(1+\delta a(y)),`$
$`n(t,y)`$ $`=`$ $`e^{|y|m_0b_0}(1+\delta a(y)),`$
$`b`$ $`=`$ $`b_0,`$ (6.41)
it is possible to derive the Friedmann equation
$$H^2=\frac{\kappa ^2m_0}{3(1\mathrm{\Omega }_0^2)}(\delta V_p+\delta V_T\mathrm{\Omega }_0^4),$$
(6.42)
where $`\mathrm{\Omega }_0e^{m_0b_0/2}`$. Note that (6.42) is the standard Hubble law with correct normalization for the physically observed energy density $`\rho =\delta V_p+\delta V_T\mathrm{\Omega }_0^4`$.
The constraint between the matter on the two branes was a consequence of trying to find a static solution to the radion equations of motion without actually providing a mechanism for stabilization. Once such a mechanism is introduced the constraint vanishes as described above, and the ordinary 4-dimensional FRW behavior is recovered at low temperatures if the radion has a mass of order the weak scale. It was suggested in that matter on the hidden brane or in the bulk may be a dark matter candidate.
As we have already discussed above, it seems unlikely that the RSI scenario as presented in can provide a physically realistic cosmological model as the energy density on the TeV brane is negative. The RSII model, having non-compact extra dimension, has greater success as a cosmological model in that it correctly reproduces the conventional cosmology on the visible brane (see e.g. ). Other variations of both RSI and RSII and alternative brane world models have also produced correct cosmological behaviors of our universe. The reader is referred to the review for a detailed summary of work on RS cosmology.
Important problems and challenges which need to be explained in brane world scenarios include the stabilization of the radius of the extra dimension and the radion field \- , inflation -, incorporation into supergravity models -, string theory and the AdS/CFT correspondence -. In particular see the review and the references therein. For more on the cosmological constant and brane worlds see and -. For early versions of brane world scenarios see -. Experimental predictions are discussed in e.g. -. Cosmologies of brane world scenarios are analyzed in -.
### 6.5 Supersymmetry
We will have only a few comments in this section, as the work in this area is still too new to review. There have been a number of attempts to include supersymmetry into the RS and brane world scenarios -. Supersymmetry may play an important role in many aspects of brane world models such as fine-tuning between bulk and brane cosmological constants and the stabilization of the fifth dimension (BPS vacua are stable against perturbations). Furthermore, supersymmetry and supergravity are critical aspects of string theory and hence it should be expected that they will play an integral role in string theory realizations of brane world scenarios.
Although there was legitimate concern that brane world models may be impossible to realize as BPS or non-BPS configurations of a supersymmetric theory , recent work has found a way to circumvent these no-go theorems (see, e.g. ). In the authors obtain the original Randall-Sundrum configuration from type IIB supergravity. This is achieved by considering a solution to the $`D=10`$ type IIB supergravity equations which has a $`5`$D interpretation. Note however that this is not fully a $`D=5`$ solution as it requires the $`S^5`$ massive Kaluza-Klein breathing mode. Breathing modes of sphere reductions are often useful in supporting domain walls . In this model it is possible to recover the single brane RSII model by pushing the hidden brane off to the Cauchy horizon of AdS. Another pleasing feature of this model is that the D3-brane configuration is dynamically stable.
Another interesting work provides a supersymmetric version of the minimal RS model in which the branes are singular .
Note that not all scenarios involving large extra dimensions rely on supersymmetry, such as the ADD model described in . The ADD scenario is not without its own troubles however as it has light KK gravitinos which could cause drastic problems with nucleosynthesis and the cosmic gamma ray background .
As an increasing number of works on the supersymmetrization of the RS model become available we will no doubt gain a better understanding of how this configuration should be assimilated into models of M/superstring theory and the AdS/CFT correspondence.
## 7 Conclusions
In this review we have discussed a number of intriguing approaches to string and M-theory cosmology. While the past few years have shown a considerable increase in our understanding of M-theory, there is still plenty of room for further research.
Perhaps the greatest advances have come from the discovery of duality symmetries in the M-theory moduli space, D-branes, the AdS/CFT correspondence and the development of Matrix theory. As demonstrated in this review we have taken the first steps to incorporate this new knowledge into cosmology. M-theory provides an innovative framework in which to study the early Universe and to search for alternatives to the Standard Big-Bang and Inflationary models. Conversely, cosmology is essential to our study of M-theory, since couplings and masses set by the vacuum state of string theory must agree with those observed in our Universe. The amalgamation of M-theory and cosmology may reveal the answers to a number of tantalizing questions and provide the tools to probe the earliest moments of creation.
## Acknowledgments
I wish to thank D. Dooling, A. Jevicki and especially R. Brandenberger and D. Lowe for many useful comments and discussions. I would also like to thank the PIMS at the University of British Columbia and the members of the string cosmology conference held there in July 2000. In particular, I am grateful to E. Akhmedov, S. Alexander, A. Ghosh, B. Greene, T. Hertog, B. Ovrut, G. Veneziano, H. Verlinde and T. Wiseman for helpful discussions. This work was supported in part by the Graduate Assistance in Areas of National Need (GAANN) doctoral fellowship program. |
warning/0003/nlin0003064.html | ar5iv | text | # 1 Introduction
## 1 Introduction
In the past few years supersymmetric quantum mechanics based on shape invariance and intertwining concepts has manifested a significant progress -. Its technique started to influence not only the traditional branches of physics such as atomic, nuclear and high energy physics, which originally stimulated its emergence , but also the classical areas of mathematical physics and the theory of differential equations. Recently, in - the detailed investigation of factorization technique has been performed for one specific form of second order differential equations (SODE) with polynomial coefficients, admitting polynomial solutions based on known Rodrigueโs formula . We choose similar initial arguments to construct explicitly a wide class of QM potentials for 1D Schrรถdinger equations admitting (after separation of asymptotic behaviour of the wave function) polynomial solutions.
Our consideration and analysis in sections 2 and 3 shares a common subject with the known Natanzon papers , but differs from them by being more general because of the fact that we work with Papperitz rather than hypergeometric equation with arbitrary positions of the singular points (in complex domain). Moreover, in addition to transformations described in , where the spectral parameters are preserved, we include into consideration the cases when we change the roles of spectral parameters in original polynomial system and in the Schrรถdinger equation (generalizing in this way the known consideration of the Coulomb case, see the discussion in the text).
One more thing we would like to mention is that Turbinerโs approach to generalized Bocher problem is very near to ours for exactly solvable case (though in the possibility of the transformation to appropriate Sturm-Liouville problem is merely mentioned rather than investigated in full details). But for quasi-solvable cases our approach differs from those in because we do not use the factorization ideas and investigate symmetry properties but concentrate on the connection of the appropriate polynomials with the corresponding family of Schrรถdinger equations. It allows us to explicitly define the additional relations for the operator to be zero-grading (in terms of ) and gives the way of regular construction of quasi-solvable potentials (all inside some definite family) with an arbitrary chosen number of algebraically constructed eigenstates.
The last remark we should make is that we do not intend to perform the comparison of the proposed approach with all other known methods of the construction of exactly solvable quantum potentials as this should be definitely the topic of a review paper rather than original research paper. But, in our opinion, the proposed approach will definitely influence the reviewing of the results obtained in other ways (especially for quasi-solvable cases) and further establishment of its inherent relations with other methods.
The paper is organized as follows. In the Section 2 we start with the special case leading to polynomial solutions, namely the polynomial family introduced in -. Section 3 is devoted to the explicit construction of the Schrรถdinger equation corresponding to that polynomial family, and to the presentation of some classification scheme as well as the discussion of its relation to the known solvable cases. In Section 3 we represent the generalization of the proposed scheme in some directions, namely the application to the partial differential equations and to higher order polynomial coefficients, and we demonstrate one non-trivial irreducible example and make some concluding remarks on the applicability of the considered approach.
## 2 One Construction of Polynomial Solutions for SODE
Practically all solvable 1D quantum problems correspond to those which can be transformed to the equation of hypergeometric type, and that gives eigenvalues of bound states by the requirement of finite hypergeometric series, thus being a polynomial of a given order . We start the consideration from the case leading precisely to that known situation, though from a little bit another point of view. Recently - it was shown that the eigenvalue problem for the operator of the form
$`\widehat{}={\displaystyle \frac{1}{W(x)}}{\displaystyle \frac{d}{dx}}\left(A(x)W(x){\displaystyle \frac{d}{dx}}\right)`$ (1)
leads to polynomial solutions with special requirements for the functions $`A,W`$. Namely, if we choose $`A(x)`$ as a polynomial of at most a second order, let us define,
$$A(x)=a_0+a_1x+a_2x^2,$$
(2)
and $`W(x)`$ as a non-negative function such that $`\frac{1}{W(x)}\frac{d}{dx}\left(A(x)W(x)\right)`$ is at most a first order polynomial
$$B(x)=b_0+b_1x,$$
(3)
then we can construct an orthogonal polynomial family being a solution for the eigenvalue problem, namely for the operator $`\widehat{}`$,
$`{\displaystyle \frac{1}{W(x)}}{\displaystyle \frac{d}{dx}}\left(A(x)W(x){\displaystyle \frac{d\mathrm{\Phi }_n(x)}{dx}}\right)+\gamma _n\mathrm{\Phi }_n(x)=0.`$ (4)
The polynomials given by the classical Rodrigueโs formula
$`\mathrm{\Phi }_n(x)={\displaystyle \frac{a_n}{W(x)}}\left({\displaystyle \frac{d}{dx}}\right)^n\left(A^n(x)W(x)\right)`$ (5)
are orthogonal with respect to the weight function W(x) on the interval $`(a,b)`$, chosen such that the following conditions hold
$$A(a)W(a)=A(b)W(b)=0.$$
(6)
This can be shown by classical consideration as exposed e.g. in . In this way we can choose the interval as inside the roots of the polynomial $`A(x)`$, if the latter one possesses the real roots, or the roots of $`W(x)`$, including infinity points, in the case of the functions tending to zero at infinity, or some combination of both alternatives. The eigenvalues $`\gamma _n`$ turn out to be given as -
$`\gamma _n=n\left({\displaystyle \frac{(A(x)W(x))^{}}{W(x)}}\right)^{}{\displaystyle \frac{n(n1)}{2}}A^{\prime \prime }(x),`$ (7)
which in case (2) and (3) is equal to expression (11).
At that point, having enough information, we can write and solve the equation for $`W(x)`$, namely
$`{\displaystyle \frac{d}{dx}}\left(A(x)W(x)\right)B(x)W(x)=0`$ (8)
As the equation is linear ODE of the first order its solution has the form, explicitly
$`W(x)={\displaystyle \frac{C}{A(x)}}\mathrm{exp}\left\{{\displaystyle \frac{B(x)dx}{A(x)}}\right\}={\displaystyle \frac{C\mathrm{exp}\left\{\frac{b_0+b_1x}{a_0+a_1x+a_2x^2}๐x\right\}}{a_0+a_1x+a_2x^2}},`$ (9)
and, of course, the integral we wrote is easy to calculate
$`{\displaystyle \frac{b_0+b_1x}{a_0+a_1x+a_2x^2}๐x}={\displaystyle \frac{b_1\mathrm{log}(a_0+a_1x+a_2x^2)}{2a_2}}`$ (10)
$`{\displaystyle \frac{\left(2a_2b_0+a_1b_1\right)\mathrm{arctan}(\frac{a_1+2a_2x}{\sqrt{a_{1}^{}{}_{}{}^{2}+4a_0a_2}})}{a_2\sqrt{a_{1}^{}{}_{}{}^{2}+4a_0a_2}}},`$
but for our purposes it is convenient to use it in the form we represent $`W(x)`$ in equation (9).
It is evident from the direct substitution that in the case (2) and (3) the eigenvalues $`\gamma _n`$ are given explicitly by
$`\gamma _n=n(b_1+a_2(n1)),`$ (11)
while the equation for the polynomials becomes
$`(a_0+a_1x+a_2x^2)\mathrm{\Phi }_n^{\prime \prime }(x)+(b_0+b_1x)\mathrm{\Phi }_n^{}(x)`$ (12)
$`n(b_1+a_2(n1))\mathrm{\Phi }_n(x)=0`$
The equation (12) is SODE with 3 singular points at the roots of $`A(x)`$ (we denote them as $`x_1,x_2`$) and infinity. It is easy to check that if the roots of $`A(x)`$ are different the equation is of hypergeometric type (generally speaking Papperitz equation ), whilst it is confluent hypergeometric when the roots are coincident.
The question naturally appearing is what sort of quantum mechanical problems could be associated with the polynomial family we described. To answer it, there is a natural way, namely we can try to implement adjusted pair of coordinate transformation and similarity transformation for the equation (12) in such a way, as to obtain the constant coefficient at second derivative and zero coefficient at first derivative. The resulting equation will be of Schrรถdinger type. Let us perform this program.
## 3 Transformation to the Schrรถdinger equation
Let us forget for a while that $`A(x)`$ and $`B(x)`$ are polynomials. We have the equation
$`A_2(x)y^{\prime \prime }(x)+A_1(x)y^{}(x)+\epsilon y(x)=0`$ (13)
with arbitrary coefficient functions $`A_1(x),A_2(x)`$. First, we make the variable change
$`x=F(u),`$
$`{\displaystyle \frac{d}{dx}}={\displaystyle \frac{1}{F^{}(u)}}{\displaystyle \frac{d}{du}},`$
$`{\displaystyle \frac{d^2}{dx^2}}={\displaystyle \frac{1}{F^2(u)}}{\displaystyle \frac{d^2}{du^2}}{\displaystyle \frac{F^{\prime \prime }(u)}{F^3(u)}}{\displaystyle \frac{d}{du}},`$ (14)
and choose the transformation in a form allowing to introduce some yet undefined but prescribed function of new coordinate $`\omega (u)`$ (which we could define later for the sake of the most convenient choice)
$$\omega ^2(u)[F^{}(u)]^2=A_2(x)$$
(15)
to get
$`\omega ^2(u)y^{\prime \prime }(u)+y^{}(u)\omega (u)\left({\displaystyle \frac{2A_1(F(u))A_2^{}(F(u))}{\sqrt{A_2(F(u))}}}+\omega ^{}(u)\right)`$ (16)
$`+\gamma y(u)=0`$
where prime means the differentiation with respect to the functionโs argument. Now we implement the similarity transformation $`Y(u)=\mathrm{exp}(\chi (u))y(u)`$ and choose the function $`\chi (u)`$ in such a way as to kill the term with the first derivative, so that we must have
$`\chi {}_{}{}^{}(u)={\displaystyle \frac{1}{2\omega (u)}}({\displaystyle \frac{A_{2}^{}{}_{}{}^{}(F(u))}{2\sqrt{A_2(F(u))}}}{\displaystyle \frac{A_1(F(u))}{\sqrt{A_2(F(u))}}}\omega ^{}(u)).`$ (17)
Then, the equation is transformed into
$`\omega ^2(u)Y^{\prime \prime }(u)+Y(u)(\gamma {\displaystyle \frac{A_1(F(u))^2}{4A_2(F(u))}}+{\displaystyle \frac{A_1(F(u))A_{2}^{}{}_{}{}^{}(F(u))}{2A_2(F(u))}}`$ (18)
$`{\displaystyle \frac{3A_{2}^{}{}_{}{}^{}(F(u))^2}{16A_2(F(u))}}+{\displaystyle \frac{\omega ^{}(u)^2}{4}}+{\displaystyle \frac{A_{2}^{}{}_{}{}^{\prime \prime }(F(u))}{4}}{\displaystyle \frac{A_{1}^{}{}_{}{}^{}(F(u))}{2}}{\displaystyle \frac{\omega (u)\omega ^{\prime \prime }(u)}{2}})`$
which can be considered as Schrรถdinger type equation if we manage to identify and to separate some free constant parameter in it at $`Y(u)`$ (playing the role of โenergyโ), after the division of both terms by $`\omega ^2(u)`$. As we see in the trivial choice of $`\omega (u)=1`$ we simply get the Schrรถdinger equation with the energy $`\gamma `$ and the potential $`V(u)`$ given by
$`V(u)=+{\displaystyle \frac{4A_1(F(u))^28A_1(F(u))A_2^{}(F(u))3A_2^{}(F(u))^2}{16A_2(F(u))}}`$ (19)
$`{\displaystyle \frac{A_1^{}(F(u))}{2}}{\displaystyle \frac{A_2^{\prime \prime }(F(u))}{4}}.`$
One known case when we have to choose $`\omega (x)`$ not equal to unity is the Coulomb potential as we will see below. Now we can use the fact that both $`A_2(x),A_1(x)`$ are polynomials, choosing them in accordance with (2,3) as $`A_2(x)=a_2x^2+a_1x+a_0`$ and $`A_1(x)=b_1x+b_0`$ to obtain the equation
$`Y^{\prime \prime }(u)+Y(u)[{\displaystyle \frac{a_2b_1+2\gamma }{2\omega ^2(u)}}+{\displaystyle \frac{\omega ^{}(u)^2}{4\omega ^2(u)}}{\displaystyle \frac{\omega ^{\prime \prime }(u)}{2\omega (u)}}`$
$`{\displaystyle \frac{3\left(a_1+2a_2F(u)\right)^2}{16\omega ^2(u)\left(a_0+a_1F(u)+a_2F(u)^2\right)}}`$ (20)
$`+{\displaystyle \frac{\left(a_1+2a_2F(u)\right)\left(b_0+b_1F(u)\right)}{2\omega ^2(u)\left(a_0+a_1F(u)+a_2F(u)^2\right)}}`$
$`{\displaystyle \frac{\left(b_0+b_1F(u)\right)^2}{4\omega ^2(u)\left(a_0+a_1F(u)+a_2F(u)^2\right)}}]=0`$
Now, we start to analyze first systematically the simplest case putting function $`\omega (u)`$ to be unity. Then, for the potential of the Schrรถdinger equation we have
$`V(u)={\displaystyle \frac{a_2}{2}}+{\displaystyle \frac{b_1}{2}}+{\displaystyle \frac{3\left(a_1+2a_2F(u)\right)^2}{16\left(a_0+a_1F(u)+a_2F(u)^2\right)}}+`$ (21)
$`{\displaystyle \frac{\left(a_1+2a_2F(u)\right)\left(b_0+b_1F(u)\right)}{2\left(a_0+a_1F(u)+a_2F(u)^2\right)}}+{\displaystyle \frac{\left(b_0+b_1F(u)\right)^2}{4\left(a_0+a_1F(u)+a_2F(u)^2\right)}}.`$
We can explicitly find the dependence $`x=F(u)`$ by solving the equation (15). Then, taking the inverse function, we get for $`x`$
$`x=F(u)=\{\begin{array}{cc}\frac{a_1\mathrm{sinh}(\frac{\sqrt{a_2}u}{2})^2+\sqrt{a_0a_2}\mathrm{sinh}(\sqrt{a_2}u)}{a_2},& a_20,๐0\\ \frac{a_1+(a_1+2a_2)e^{\sqrt{a_2}u}}{2a_2},& a_20,๐=0\\ \sqrt{a_0}u+\frac{a_1u^2}{4},& a_2=0,a_10\\ \frac{1}{\sqrt{a_0}}u,& a_2=0,a_1=0,a_00\end{array}`$
where $`๐=a_1^24a_0a_2`$ is the discriminant for $`A_2(x)`$.
The explicit expression for the quantum potential can be obtained after the substitution of (3) into expression (21) but for the general case the resulting formula becomes too complicated. Specifying the values of the parameters $`a_i,b_i`$ it is possible to show that almost all known solvable cases in quantum mechanics except the Coulomb potential, are inside the potential family we constructed. The latter one will be analyzed later when we try to construct new potentials by choosing a nontrivial $`\omega (u)`$ function. So let us first investigate the case $`\omega (u)=1`$.
We shall classify the different cases by the roots of the polynomial $`A_2(x)`$. We have two topologically different cases when $`A_2(x)`$ is not degenerate, namely when $`A_2(x)`$ has two different roots and the discriminant $`๐0`$, and the case of one root with degeneracy 2, $`๐=0`$. As one can easily see in the space of the parameters $`a_i`$ the first case fills the region inside and outside the conical surface for which the equation is $`๐=0`$. So we call the appropriate cases regular and irregular (for $`๐=0`$), respectively. We shall refer to them sometimes as Jacobi and Morse cases, based on the name of the appropriate polynomials (for the first one) and solutions (second one).
The additional cases appear in (3) as a result of the degeneracy of the polynomial $`A_2(x)`$, it could be of the first order (we call this case Laguerreโs case; $`a_2=0`$) and of the zero order (further referred to as Hermite case; $`a_2=a_1=0`$).
The first remark we would like to make is that, our polynomial family has five parameters whereas spectra depend on top power coefficients of $`A_2(x),A_1(x)`$ only, namely $`a_2,b_1`$ (see equation (11)), and so we have the evident freedom of choosing some parameters without loosing characteristic features of the problem. Obviously, we can change parameters $`b_0,a_1`$ simply by the trivial change of the origin in $`x`$ variable and the scale on it. Then, if we choose definite values of $`a_2,b_1`$ (one of them could be considered as a scale for energy and could be chosen e.g. as unity) we obtain a two parametric family of polynomials and one parametric family with full isospectrality property. The variation of the parameter $`a_0`$ then will lead to different non-trivial cases we mentioned.
So, we start with the regular case $`๐0`$. Then the change of variables $`xu`$ is given by the top line in (3), and the orthogonal polynomials have the weight function given by
$`W(x)={\displaystyle \frac{\mathrm{exp}\left\{\frac{\left(2a_2b_0a_1b_1\right)}{a_2๐}\mathrm{arctan}(\frac{a_1+2a_2x}{๐})\right\}}{(a_0+a_1x+a_2x^2)^{1\frac{b_1}{2a_2}}}}`$ (22)
As this family of polynomials has no commonly used name we will refer to it as to generalized Jacobi polynomials, ordinary Jacobi case corresponds to $`a_0=1,a_1=0,a_2=1`$,$`b_0+b_1=2p,b_0b_1=2q`$ corresponding to symmetrically chosen real roots at $`\pm 1`$ and the interval of orthogonality $`[1,1]`$.
The quantum potential has the following general form ($`z=\sqrt{a_2}u`$)
$`V(u)={\displaystyle \frac{A+B\mathrm{sinh}z+C\mathrm{cosh}z+D\mathrm{sinh}2z+E\mathrm{cosh}2z}{(2\sqrt{a_0a_2}\mathrm{cosh}z+a_1\mathrm{sinh}z))^2}}`$ (23)
with the coefficients $`A,B,C,D,E`$ expressed through the original ones as
$`A=`$ $`a_{2}^{}{}_{}{}^{2}\left(5a_{1}^{}{}_{}{}^{2}20a_0a_2+8b_{0}^{}{}_{}{}^{2}\right)+\left(3a_{1}^{}{}_{}{}^{2}4a_0a_2\right)b_{1}^{}{}_{}{}^{2}`$
$`+\mathrm{\hspace{0.33em}2}a_2\left(3a_{1}^{}{}_{}{}^{2}+12a_0a_24a_1b_0\right)b_1,`$
$`B=`$ $`8\sqrt{a_0a_2}\left(b_12a_2\right)\left(2a_2b_0a_1b_1\right),`$
$`C=`$ $`4a_1\left(2a_2b_1\right)\left(2a_2b_0a_1b_1\right),`$ (24)
$`D=`$ $`4a_1\sqrt{a_0a_2}\left(a_2b_1\right)^2,`$
$`E=`$ $`\left(a_{1}^{}{}_{}{}^{2}+4a_0a_2\right)\left(a_2b_1\right)^2.`$
It is straightforward to see that the potential family (23) includes Pรถschl-Teller potentials (both ordinary and modified), Scarf-like potentials, Rosen-Morse and Manning-Rosen potentials at appropriate choice of the parameters.
In singular case $`D=0`$ it is more convenient to introduce other parameters rather than $`a_i`$, namely $`a_0=\alpha ^2,a_2=\beta ^2,a_1=2\alpha \beta `$, automatically satisfying the degeneracy condition, and then the weight function becomes
$`W(x)=\left(\alpha +\beta x\right)^{2+b_1}\mathrm{exp}\left\{{\displaystyle \frac{b_1\alpha b_0\beta }{\beta ^2(\alpha +\beta x)}}\right\},`$ (25)
and the potential reads through newly introduced coefficient $`A,B,C`$
$`V(u)=A+Be^{\beta u}+Ce^{2\beta u},`$
$`A={\displaystyle \frac{\left(b_1\beta ^2\right)^2}{4\beta ^2}},`$
$`B={\displaystyle \frac{\left(b_1\alpha b_0\beta \right)\left(b_12\beta ^2\right)}{2\alpha \beta ^2}}`$ (26)
$`C={\displaystyle \frac{\left(b_1\alpha b_0\beta \right)^2}{4\alpha ^2\beta ^2}}.`$
This evidently corresponds to the Morse class of potentials .
The case when $`A_2(x)`$ becomes the first order polynomial ($`a_2=0`$), gives for the weight the following formula
$`W(x)=e^{\frac{b_1x}{a_1}}\left(a_0+a_1x\right)^{1+\frac{b_0}{a_1}\frac{a_0b_1}{a_1^2}},`$ (27)
and for the potential
$`V(u)=A+Bu^2+Cu^2,`$ (28)
with the new parameters $`A,B,C`$ (not to be confused with those obtained above) expressed through the old ones as
$`\begin{array}{cc}A& =\frac{b_1(a_1b_0a_0b_1)}{2a_1^2},\\ B& =\frac{(a_1^22a_1b_0+2a_0b_1)(3a_1^22a_1b_0+2a_0b_1)}{4a_1^4},\\ C& =\frac{b_1^2}{16}.\end{array}`$
The resulting potentials, as we see, are the combination of harmonic oscillator (HO) potential plus centrifugal potential $`B/u^2`$. And for the sake of completeness it is worthwhile to mention, that the case when $`A_2(x)`$ is a constant ($`a_2=a_1=0`$) corresponds to ordinary HO case with the oscillator position shifted by $`b_0/b_1`$.
Before going further, let us construct the explicit representation for the wave functions of the appropriate Schrรถdinger equation and let us discuss the bound states within this approach. As we made two subsequent transformations to obtain the Schrรถdinger equation, the solution in terms of polynomials has the form
$`Y(u)=\mathrm{exp}\left\{\chi (u)\right\}P_n(F(u))`$ (29)
where $`\chi (u)`$ is given by (17) and $`F(u)`$ is given by (15). The energy corresponding to this eigenfunction turns out to be the sum of $`\gamma _n`$ ( see equation (11) ) and some constant factor depending upon the parameters $`a_i,b_i`$ and leading to the shift of the energyโs origin. As the family of orthogonal polynomials has infinity and countable number of members, though the class of the potentials includes not only those which grow indefinitely at infinity (supporting bound states only), but also such with a finite number of levels (and finite ionization energy), we have to understand what is the condition for a bound state in the system. Indeed, this is very simple in the discussed case, the function $`W(u)`$ gives the asymptotic behaviour of the ground state wave function and as the point transformation $`x=F(u)`$ could be non-trivial, the resulting high order polynomials $`P_n(F(u))`$ can have growing behaviour at infinity which might more than compensate that of $`W(u)`$ and thus makes $`Y(u)`$โs norm infinite. Therefore, the condition for the bound states is simply the condition of a finite norm of $`Y(u)`$. Of course, an interesting question appears whether the non-normalizable solutions of polynomial type correspond to some physically significant features of the system, e.g. to quasi-bound states (long-lived localized states) embedded in the continuum, but we will not discuss it here.
Now, we can consider other possible choices of the function $`\omega (u)`$. This is stimulated by the known sequence of transformations for the Coulomb problem , where the first step is the change of scale in a way to put โenergyโ parameter into potential function with subsequent transformation of the original Schrรถdinger equation into hypergeometric equation.
As we can see from the equation (3) the problem is pure algebraic and there are several ways to try to obtain the free parameter which could be interpreted as energy. The first one is to choose $`\omega (u)`$ to satisfy the equations either $`\omega ^{}(u)/\omega (u)=k`$ or $`\omega ^{\prime \prime }(u)/\omega (u)=k`$, that will produce constant factor due to fractions including the appropriate ratios in equation (3). The second one takes place when $`\omega ^2(u)=A_2(F(u))^k`$, $`k>0`$, that could lead to free coefficient in potential due to cancellation of some denominators in (3). We shall not pursue these cases any further, but treat the most important case instead. Namely, if the order of polynomial $`A_2(x)`$ is less than two, then new possible cases also appear, as we will see, for $`k=1`$, which turns out to be precisely the Coulomb case that we discuss below. The last and more special case is realized when $`A_2(x)`$ has different real roots and the coefficients $`b_0,b_1`$ are chosen in such a way as to construct common divisor (of the first order) for both numerators and denominators in potential. In this case choosing $`\omega (u)`$ in the form of $`\omega (u)=(F(u)x_1)`$ we also obtain free coefficient in potential. We will not consider all the above mentioned cases in detail here but restrict ourselves to one specific choice of $`A_2(x)`$, stimulated by the Coulomb problem.
Let us assume that $`A_2(x)=x`$ and we will choose the $`\omega (u)`$ in the form $`\omega ^2(u)=F(u)^k`$. Then, if we choose $`k=1`$, it is easy to see from the equation (15) that the point canonical transformation turns out to be identity, and we get the standard Coulomb case
$`Y^{\prime \prime }(u)+\left[{\displaystyle \frac{b_1^2}{4}}+{\displaystyle \frac{b_0}{2u^2}}{\displaystyle \frac{b_0^2}{4u^2}}{\displaystyle \frac{b_0b_1}{2u}}+{\displaystyle \frac{\gamma }{u}}\right]Y(u)=0.`$ (30)
The case $`k=1`$ is also a special case here, so we have the following expression for the variable change $`x=e^u`$ and the Schrรถdinger equation takes the form
$`Y^{\prime \prime }(u)+[{\displaystyle \frac{(b_01)^2}{4}}{\displaystyle \frac{\left(b_0b_12\gamma \right)}{2}}e^u{\displaystyle \frac{b_1^2}{4}}e^{2u},]Y(u)=0`$ (31)
which is the case of the quantum Morse potential.
For $`k\pm 1`$ we obtain after integration of the equation (15) the following expression for the point transformation,
$`x=\left[{\displaystyle \frac{(k1)(uC_1)}{2}}\right]^{\frac{2}{1k}}.`$ (32)
The substitution of the last expression leads to a fairly complicated form of the Schrรถdinger equation for $`Y(u)`$, but as one can show, there are no more cases except those we mentioned, where it is possible to obtain a free parameter in the role of โenergyโ.
The successful implementation of the construction of solvable Schrรถdinger potentials, as we already have seen, was due to the evident existence of polynomial solution of the equation (12). It is possible to find the generalization in more complicated cases, which is the topic of the following section.
## 4 Some generalizations of the approach
As we saw, the main feature of the system considered was that the second order differential operator $`\widehat{}`$ (see equations (1) and (12)) preserves the linear subspace $`_n`$ of the polynomials of order $`n`$ for all $`n`$. It was due to the special adjustment of the orders of polynomial coefficients with the order of appropriate derivatives. This idea, of course, can be implemented not only for a special case of SODE and polynomial coefficients up to the second order, but in much more general case of linear PDEs. Indeed, we can construct the following general form of the $`n`$-th order linear differential operator $`\widehat{}`$ with the same property, so that it preserves the space when acting in the space of the polynomials of $`m`$ variables $`\stackrel{}{x}=\{x_1,\mathrm{}x_m\}`$. The general form of the appropriate linear PDEs reads
$`\widehat{}Y(\stackrel{}{x})={\displaystyle \underset{j=0}{\overset{n}{}}}P_{j+N}(\stackrel{}{x})_{\kappa _j}Y(\stackrel{}{x})=\mathrm{\hspace{0.17em}0},`$ (33)
where $`N`$ is some non-positive integer number (when $`N<0`$ we have degenerate cases, see analogous discussion on SODE in the previous section). We introduce multi-index $`\kappa _j`$ as $`j`$-th order partial derivative over arbitrary combinations of variables in a standard way, by
$`_{\kappa _j}={\displaystyle \frac{^{i_1}}{x_1^{i_1}}}\mathrm{}{\displaystyle \frac{^{i_m}}{x_m^{i_m}}},j=i_1+\mathrm{}i_m.`$ (34)
We also say that the weight of $`\kappa _j`$ equals $`j`$ and write it as $`\mathrm{\#}\kappa _j=j`$. It is worthwhile to point out that we even do not need to demand the commutativity of the derivatives, so that the same consideration could be applied for non-commutative case (quantum groups and quantum algebras see, e.g. ). Moreover, we can consider the operators which do not preserve such finite dimensional spaces, but map one into another (with higher dimension). The latter is just the permission for $`N`$ to be positive.
In the case of standard partial derivatives, when we are looking for the solution in terms of $`k`$-th order polynomials in the ring $`F[x_1,x_2,..,x_m]`$, every term in the sum in (33) maps the argument into the space spanned by the monomials $`x^{\kappa _{N+k}}=x_1^{i_1}x_2^{i_2}\mathrm{}x_{m}^{}{}_{}{}^{i_m},\{i_1+\mathrm{}i_m=N+k\}`$. The latter space is a finite dimensional vector space and a direct sum of the spaces of symmetric homogeneous polynomials corresponding to different permutations of indices for the monomials written above. We denote the space spanned by the definite monomials of order $`k`$ as $`{}_{}{}^{(k)}๐ฏ_{[i_1\mathrm{}i_m]}^{}`$). Then we can write down the expression for the dimension for the image space for operator action
$`dim_k^N={\displaystyle \underset{j=0}{\overset{k+N}{}}}{\displaystyle \underset{\stackrel{i_1,\mathrm{},i_m=0}{i_1+\mathrm{}+i_m=m}}{\overset{m}{}}}dim^{(j)}๐ฏ_{[i_1\mathrm{}i_m]}.`$ (35)
Now, the construction of the k-th order polynomial solution
$`Y_k(\stackrel{}{x})={\displaystyle \underset{j=0}{\overset{k}{}}}C_{\kappa _j}x^{\kappa _j}`$ (36)
leads simply to the linear algebraic problem for non-trivial solution for the coefficients $`C_{\kappa _j}`$.
At this point two different cases are possible. The first one is realized if $`N0`$, that is the maximal order of polynomial coefficient is less or equal to the PDEโs order. In this case we can always satisfy the system of equations because the number of linear homogeneous equations for $`C_{\kappa _j}`$ is precisely equal to $`dim_k^0`$. Then, the non-triviality condition is the condition of zero determinant for corresponding matrix obtained from equating all coefficients at monomials of type $`x^{\kappa _i}`$ to zero, and this gives us spectral parameter for the polynomial family, namely the quantization condition imposed on the coefficient $`P_0(x)`$. Then, considering the problem over the field of complex numbers, we can always construct a polynomial family in this case for some quantized value of the coefficient $`P_0(x)`$. In the contrary, when we have the condition $`N>0`$, we are still obliged to fulfill $`dim_k^N`$ conditions but only for $`dim_k^0`$ coefficients $`C_{\kappa _j}`$. The resulting system becomes overcomplete which simply means that we can construct some separate polynomial solutions of the equation (33) for only a few levels, maybe even one, that is for definite choice of $`n`$ and, additionally for special values of some of the coefficients in the coefficient polynomials.
It is very interesting to mention here that in the 1-D case considered in the previous section for SODE, the appropriate matrix turns out to be upper-three-diagonal, with additional relationship between elements, so that its determinant for $`j`$-th order system has the following form (preserving notions for coefficient of $`A_2(x),A_1(x)`$)
$`\left(\begin{array}{ccccc}& \gamma & b_0\mathrm{\hspace{0.33em}\hspace{0.33em}\; \hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\; 2}a_0& \mathrm{}& 0\\ & 0& \gamma +b_1\mathrm{\hspace{0.33em}\hspace{0.33em}\; 2}b_0& \mathrm{}& 0\\ & \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ & 0& \mathrm{}& (j1)b_0+(j1)(j2)a_1& j(j1)a_0\\ & 0& \mathrm{}& \gamma +(j1)b_1+(j1)(j2)a_2& jb_0+j(j1)a_1\\ & 0& \mathrm{}& 0& \gamma +jb_1+j(j1)a_2\end{array}\right)`$
that leads to one degenerate eigenvalue (rather than $`j+1`$) for a given $`j`$-th order polynomial.
What realizations of the scheme described above could be successfully used for the construction of the solutions for Schrรถdinger equation except those which we demonstrated in Section 2? There are three evident but not easy ways. The first and the easiest one is the consideration of 1D problems with higher order polynomials, and we intend to demonstrate this in one example, putting away its full description as a subject of separate publication. We consider the construction of the polynomial solutions starting from the third order polynomials to give the representation of the problems emerging there. Let us have the equation of the form
$`x^3y^{\prime \prime }(x)+\alpha (x^21)y(x)+(\beta x+\gamma )y(x)=0.`$ (37)
Then, in the same manner as we did in Section 2 we map the equation to the Schrรถdinger equation via two subsequent point transformations $`x=4/u^2`$, and the gauge transformation
$`\chi (u)=C_1+{\displaystyle \frac{\alpha u^4}{64}}+{\displaystyle \frac{32\alpha }{2}}\mathrm{log}u.`$ (38)
The resulting equation has the form
$`Y^{\prime \prime }(u)+\left(\gamma V(u)\right)Y(u)=0,`$ (39)
$`V(u)={\displaystyle \frac{(3\alpha \alpha ^2)u^2}{8}}+{\displaystyle \frac{u^6\alpha ^2}{256}}+{\displaystyle \frac{4\alpha ^28\alpha 16\beta +3}{4u^2}}.`$ (40)
The solution will be given by the formula (29), but now we have to construct the polynomials in a non-trivial way, because the Rodrigueโs formula is no more applicable. So we start to look directly for the polynomial solution of the equation (37). Let us define $`n`$-th order polynomial as $`p_n(x)=_{i=0}^nc_ix^i`$. When we substitute this anzatz into equation (37) and equate to zero the coefficients at all orders of independent variable $`x`$, we obtain $`n+1`$ equations for $`n`$ coefficients $`c_i,i=0\mathrm{}n1,\gamma `$ (the last coefficient $`c_n`$ should be chosen in order to satisfy the standardization condition). Therefore the system is overcomplete and for a non-trivial solution we must specify some additional coefficient in a unique way or to impose one additional condition on some parameters of the system. In our case we have the only choice to add $`\beta `$ to the list of coefficients to be found. Let us find then the solution for example for $`n=1,2`$ explicitly starting from the case $`n=1`$.
Solving the set of equations for $`c_0,\beta ,\gamma `$, two solutions $`c_0^\pm =\pm \mathrm{\hspace{0.17em}1},\beta =\alpha ,\gamma ^\pm =\alpha `$ can be constructed that giving for $`Y_1(u)`$
$`Y_1^\pm (u)=e^{\frac{\alpha u^4}{64}}\left(1+{\displaystyle \frac{4}{u^2}}\right)u^{\frac{32\alpha }{2}},(\gamma =\alpha ).`$ (41)
We have to impose the finiteness condition for the norm of the solution, so that in our case $`\alpha <1/4`$. The last requirement follows from simultaneous demand on proper behaviour at infinity that leads to $`\alpha <0`$ and integrability at $`u=0`$. But as one can see, in the interval $`1/2<\alpha 1/4`$ the potential becomes repulsive as $`u0`$. The last property means that we have to impose the boundary condition $`Y(u)|_{u=0}=0`$ that leads to the restriction $`\alpha 1/2`$ on the admittable region for the variation of $`\alpha `$.
Then, both solutions will be, as one can see, bound states of the system and the solution for larger $`\gamma `$ will have zero not only at $`u=0`$ but also at $`u=2`$, which evidently corresponds to the first excited state, whereas the first one corresponds to the ground state.
In a similar way, for $`n=2`$ we obtain the condition $`\beta =2\left(1+\alpha \right)`$ and the following three solutions for $`c_0,c_1,\gamma `$
$`c_0^{(1)}={\displaystyle \frac{\alpha }{\alpha +1}},c_1^{(1)}=0\gamma ^{(1)}=0,`$ (42)
$`c_0^{(\pm )}={\displaystyle \frac{\alpha }{\alpha +2}},c_1^{(\pm )}=\pm {\displaystyle \frac{\sqrt{2\alpha \left(2\alpha +3\right)}}{\alpha +2}},\gamma ^{(\pm )}=\pm 2\sqrt{2\alpha \left(2\alpha +3\right)}.`$
Then, the appropriate eigenfunctions are given by
$`Y^{(2)}(u)=e^{\frac{\alpha u^4}{64}}u^{\frac{3}{2}\alpha }\left({\displaystyle \frac{16}{u^4}}{\displaystyle \frac{\alpha }{\alpha +1}}\right),`$ (43)
$`Y^{(\pm )}(u)=e^{\frac{\alpha u^4}{64}}u^{\frac{3}{2}\alpha }\left({\displaystyle \frac{16}{u^4}}+{\displaystyle \frac{\alpha }{\alpha +2}}\pm {\displaystyle \frac{4\sqrt{2\alpha \left(2\alpha +3\right)}}{(\alpha +2)u^2}}\right).`$
The similar consideration as we performed for $`n=1`$ shows that admittable region for the parameter $`\alpha `$ now is given by $`\alpha 5/2`$. Then the constructed eigenstates represent the ground and the first two excited states for the potential
$`V(u)={\displaystyle \frac{\alpha ^2u^6}{256}}{\displaystyle \frac{\alpha \left(\alpha 3\right)u^2}{8}}+{\displaystyle \frac{4\alpha ^2+24\alpha +35}{4u^2}}.`$ (44)
The remarkable feature of the example we considered was that we construct some eigenstates corresponding to a given potential using polynomial anzatz of a given order. The solution for those cases corresponding to the ground state and to the excited states was given by the number of the roots of the
polynomials of a given order. This is in contradistinction with the standard solvable situation where polynomialโs order is equal to the quantum number because for the classical polynomial family, the $`n`$-th order polynomial has precisely $`n`$ roots on an interval where the family is defined, unlike in the general case, where the real polynomial can have complex (nonreal) roots.
ยฟFrom the last we can make some useful conclusions. As we see, if the order of polynomial coefficient functions is different from the order of the equation, the polynomial solutions in general do not exist except for the special values of the parameters. In the example that has been considered, we have to put one additional condition on the parameter to construct non-trivial solution. Nevertheless, after appropriate restriction of the region for some parameters ($`\alpha `$ in the considered example) and by fixing the value for some other (expressing $`\beta `$ through the $`\alpha `$ in the discussed case) we were able to construct eigenfunctions of some low lying eigenstates
Now, we return to the discussion of other possible generalizations of the proposed approach.
The second admissible choice is to consider, in the same way as we did, the Schrรถdinger equations solvable in momentum representation. As it is easy to understand, physically interesting potentials of polynomial type with the order greater than two will correspond to higher order differential equations, so that we can e.g. ask which Schrรถdinger equations can be constructed, based on the polynomial solutions for the equation like this $`A_4(x)y^{\prime \prime \prime \prime }(x)+A_3(x)y^{\prime \prime \prime }+\mathrm{}=0`$? We stop the discussion of this possibility at this point leaving it for the future publications too.
The last and the most complicated case corresponds to the consideration of genuine higher dimensional problems and appropriate PDEs. The most difficult obstacle to be overcome there is the necessity to perform transformations from original equation for polynomials to the equation of the Schrรถdinger type. Although the theory of characteristics is applicable in this case, the resulting equation, at first glance, could be hardly interpreted in terms of the Schrรถdinger type. As for the latter one we must demand the existence of pure constant term included into coefficient function at zero-derivative (see the discussion after equation (18)). Nevertheless, this direction is of great importance for deeper insight into integrability and solvability problems in quantum mechanics.
## 5 Discussion
In summary, we demonstrated that one can construct explicit formulae for the family of the orthogonal polynomials depending on five parameters, and thus we can associate with them the family of isospectral potentials (isospectrality with respect to 3 free parameters) which include almost all known quantum mechanically exactly solvable potentials. Some generalization of the approach to higher dimensional equations (PDEs) as well as to the higher order ODEs has been proposed.
We may conclude with a little bit speculative thought. If one would put forward the requirement for the bound states of a quantum system (in analytical case) as a demand on polynomial type of the reduction of the wave function (that seems to be reasonable and evidently fulfilled for all up-to-date known 1D solvable cases), then the immediate conclusion follows that the proposed approach (with its generalizations discussed) includes all analytically solvable cases.
## Acknowledgements
The authors are grateful to the referees for drawing our attention to the papers of Natanzon and Turbiner and for stimulating the improvement of the manuscript. This research has been supported by the Ministry of Science and Technology of the Republic of Slovenia. GK acknowledges also the postdoctoral research grant of the Slovenian Science Foundation. |
warning/0003/math0003068.html | ar5iv | text | # Ricci Curvature, Minimal Volumes, and Seiberg-Witten Theory
## 1 Introduction
The aim of modern Riemannian geometry is to understand the relationship between topology and curvature. It is thus interesting and natural to consider differential-topological invariants of a smooth compact $`n`$-manifold which, by their very definition, represent quantitative obstructions to the existence of a scalar-flat (or Ricci-flat, or flat) metric on the given manifold. For example, if $`M`$ is a smooth compact $`n`$-manifold, one may define invariants
$`_s(M)`$ $`=`$ $`\underset{g}{inf}{\displaystyle _M}|s_g|^{n/2}๐\mu _g`$
$`_r(M)`$ $`=`$ $`\underset{g}{inf}{\displaystyle _M}|r|_g^{n/2}๐\mu _g`$
$`_{}(M)`$ $`=`$ $`\underset{g}{inf}{\displaystyle _M}||_g^{n/2}๐\mu _g`$
where the infima are to be taken over all metrics $`g`$ on $`M`$, where $`s`$, $`r`$, and $``$ respectively denote the scalar, Ricci, and Riemann curvatures of $`g`$, and where $`d\mu `$ denotes Riemannian volume measure. The power of $`n/2`$ is used for reason of scale invariance: any other choice would result in the zero invariant. Still, one has every right to suspect that invariants with such soft definitions will either be trivial, or else will be completely impossible to calculate in practice.
This suspicion would seem to be vindicated by a recent result of Petean , who, building upon the earlier work of Gromov-Lawson and Stolz , showed that $`_s(M)=0`$ for any simply connected $`n`$-manifold, $`n5`$. Dimension $`4`$, however, turns out to be radically different. Seiberg-Witten theory naturally leads to non-trivial lower bounds for $`_s`$ which, amazingly, are often sharp. Using this, the author has elsewhere computed $`_s(M)`$ for all complex surfaces $`M`$ with even first Betti number; it turns out that $`_s(M)`$ is positive exactly for the surfaces of general type, and for these it is given by the formula
$$_s(M)=32\pi ^2c_1^2(X),$$
where $`X`$ is the minimal model of $`M`$, in the sense of Kodaira.
One main purpose of the present article is to similarly calculate $`_r`$ for the complex surfaces. Unlike the $`_s`$, this invariant is changed by blowing up, and in Theorem 4.3 we will see that it is given by
$$_r(M)=8\pi ^2[c_1^2(X)+k]$$
where $`k`$ is the number of points of $`X`$ which must be blown up in order to obtain $`M`$.
This, however, is just one application of some new curvature estimates we will develop here. Fundamentally, the story is not about the Ricci curvature at all, but instead principally concerns the rรดle of the self-dual Weyl curvature. Recall that the $`2`$-forms on an oriented $`4`$-manifold decompose as
$$\mathrm{\Lambda }^2=\mathrm{\Lambda }^+\mathrm{\Lambda }^{},$$
where $`\mathrm{\Lambda }^\pm `$ is the $`(\pm 1)`$ eigenspace of Hodge star operator $``$. Thinking of the curvature tensor $``$ as a linear map $`\mathrm{\Lambda }^2\mathrm{\Lambda }^2`$, we thus get a decomposition
=(
+W+s12BB+W-s12 )
+W+s12BB+W-s12 {\mathcal{R}}=\left(\mbox{
\begin{tabular}[]{c|c}&\\
$W_{+}+\frac{s}{12}$&$B$\\
&\\
\cline{1-2}\cr&\\
$B^{*}$&$W_{-}+\frac{s}{12}$\\
&\\
\end{tabular}
}\right)
into irreducible pieces. Here the self-dual and anti-self-dual Weyl curvatures $`W_\pm `$ are the trace-free pieces of the appropriate blocks. The scalar curvature $`s`$ is understood to act by scalar multiplication, and $`B`$ amounts to the trace-free part $`\stackrel{}{r}=r\frac{s}{4}g`$ of the Ricci curvature, acting on anti-self-dual 2-forms by
$$\psi _{ab}\underset{ac}{\overset{}{r}}\psi _{}^{c}{}_{b}{}^{}\underset{bc}{\overset{}{r}}\psi _{}^{c}{}_{a}{}^{}.$$
The key estimate of the article is to be found in Theorem 2.3: if $`(M,g)`$ is an oriented Riemannian $`4`$-manifold, and if, for a fixed spin<sup>c</sup> structure on $`M`$, the Seiberg-Witten equations have a solution for every metric conformally related to $`g`$, then the curvature of $`g`$ satisfies
$$_M\left(\frac{2}{3}s2\sqrt{\frac{2}{3}}|W_+|\right)^2๐\mu 32\pi ^2(c_1^+)^2,$$
where $`c_1^+`$ is the self-dual part, with respect to $`g`$, of the first Chern class of the spin<sup>c</sup> structure. Information about the Ricci curvature is then indirectly extracted from this, by means of the Gauss-Bonnet-like formula
$$(2\chi +3\tau )(M)=\frac{1}{4\pi ^2}_M\left(2|W_+|^2+\frac{s^2}{24}\frac{|\stackrel{}{r}|^2}{2}\right)๐\mu ,$$
(1)
which holds for every Riemannian metric $`g`$ on $`M`$. Here $`\chi (M)`$ and $`\tau (M)`$ denote the Euler characteristic and signature of $`M`$, respectively.
The above estimate should be compared and contrasted to the analogous inequality
$$_Ms^2๐\mu 32\pi ^2(c_1^+)^2.$$
for the scalar curvature, keeping in mind that $`|s|=2\sqrt{6}|W_+|`$ for any Kรคhler metric. This seems all the more remarkable insofar as the older estimate is only saturated by constant-scalar-curvature Kรคhler metrics, while the new estimate is apparently saturated by a larger class of almost-Kรคhler metrics. This reflects an under-utilized aspect of Taubesโ construction of solutions of the Seiberg-Witten equations on symplectic manifolds , and would appear to be a promising avenue for further research.
These estimates also give one new obstructions to the existence for Einstein metrics. Recall that that a smooth Riemannian metric $`g`$ is said to be Einstein if its Ricci curvature $`r`$ is a constant multiple of the metric:
$$r=\lambda g.$$
Not every smooth compact oriented 4-manifold $`M`$ admits such a metric. Indeed, a well-known necessary condition is that $`M`$ must satisfy the Hitchin-Thorpe inequality $`2\chi (M)3|\tau (M)|`$, where again $`\chi `$ and $`\tau `$ denote the signature and Euler characteristic. Indeed, this is an immediate consequence of (1), since the Einstein condition may be rewritten as $`\stackrel{}{r}=0`$, and $`\stackrel{}{r}`$ makes the only negative contribution to the integrand. Notice, however, that one could strengthen the conclusion given an a priori lower bound for the term
$$\frac{1}{4\pi ^2}_M\left(2|W_+|^2+\frac{s^2}{24}\right)๐\mu .$$
But a lower bound is not difficult to extract from the above estimate, and this allows one to show (Corollary 3.4) that if a minimal complex surface $`X`$ is blown up $`k`$ times, the resulting complex surface
$$M=X\mathrm{\#}k\overline{}_2$$
does not admit Einstein metrics if $`k\frac{1}{3}c_1^2(X)`$. This improves previous results , where the same conclusion is reached for larger values of $`k`$.
The invariant $`_{}`$ introduced at the beginning of this introduction is often easy to compute in dimension $`4`$, because
$$_{}(M)8\pi ^2\chi (M),$$
with equality if $`M`$ admits an Einstein metric. In particular, it is trivial to read off this invariant for a compact hyperbolic $`4`$-manifold $`^4/\mathrm{\Gamma }`$ or for a complex-hyperbolic $`4`$-manifold $`_2/\mathrm{\Gamma }`$. (The complex hyperbolic plane $`_2`$ may abstractly be defined as the symmetric space $`SU(2,1)/U(2)`$, but it is typically more useful to think of it as the unit ball in $`^2`$, equipped with the Bergmann metric.) It might be tempting to assume that this tells one everything there is to know about the sectional curvatures of metrics on these spaces; but in reality, that is wide off the mark! For example, one might ask whether the standard metric, suitably normalized, has least volume among all metrics of sectional curvature $`K1`$. For real-hyperbolic manifolds the answer is affirmative, and indeed this holds in all dimensions; however, the proof, due to Besson-Courtois-Gallot , depends not only on a remarkable inequality concerning volume entropy, but also on an optimal application of Bishopโs inequality. For complex-hyperbolic manifolds, the latter breaks down, and the question therefore remains open.
While the above question cannot be settled by present means, our estimates do allow one to answer a closely related question. Notice that a $`4`$-manifold with $`K1`$ also satisfies
$$\frac{1}{2}\left(K+\frac{s}{12}\right)1.$$
Given a complex hyperbolic $`4`$-manifold, we are able to show (Theorem 5.3) that, among metrics on a complex-hyperbolic manifold satisfying this curvature constraint, an appropriate multiple of the standard metric has least volume. Similar results also hold metrics subject to the curvature constraint
$$tK+(1t)\frac{s}{12}1$$
for any constant $`t[0,1/2]`$.
## 2 Weyl Curvature Estimates
In this section, we derive new $`L^2`$ estimates for combinations of the Weyl and scalar curvatures of certain Riemannian $`4`$-manifolds. These considerably refine the estimates previously found in .
Let $`M`$ be a smooth, compact, oriented 4-manifold. Each Riemannian metric $`g`$ on $`M`$ then determines a direct sum decomposition
$$H^2(M,)=_g^+_g^{},$$
where $`_g^+`$ (respectively, $`_g^{}`$) consists of those cohomology classes for which the harmonic representative is self-dual (respectively, anti-self-dual). The non-negative integer $`b_+(M)=dim_g^+`$ is independent of $`g`$, and we will henceforth always assume it to be positive. It is thus natural to consider the set of metrics $`g`$ for which $`_g^+=H`$ for some fixed $`b_+(M)`$-dimensional subspace $`HH^2(M,)`$; such metrics will be said to be $`H`$-adapted. Assuming there is at least one $`H`$-adapted metric, we will then say that $`H`$ is a polarization of $`M`$, and call the pair $`(M,H)`$ a polarized 4-manifold . Notice that the restriction of the intersection pairing
$$:H^2(M,)\times H^2(M,)$$
to $`H`$ is then positive definite, and that $`HH^2`$ is maximal among subspaces with this property.
Let $`c`$ be a spin<sup>c</sup> structure on $`M`$. Then $`c`$ determines a Hermitian line-bundle $`LM`$ with
$$c_1(L)w_2(M)mod2,$$
and for each metric $`g`$ we also have rank-2 complex vector bundles $`๐_\pm M`$ which formally satisfy
$$๐_\pm =๐_\pm L^{1/2},$$
where $`๐_\pm `$ are the locally-defined left- and right-handed spinor bundles of $`g`$. Given a polarization $`H`$ on $`M`$, we will then use $`c_1^+`$ to denote the orthogonal projection of $`c_1(L)`$ into $`H`$ with respect to the intersection form. If $`g`$ is a particular metric with $`_g^+=H`$, we will also freely use $`c_1(L)`$ to denote the $`g`$-harmonic 2-form representing the corresponding de Rham class, and use $`c_1^+`$ to denote its self-dual part. For example, if a choice of $`H`$-compatible metric $`g`$ has already been made, the number
$$|c_1^+|:=\sqrt{(c_1^+)^2}$$
may freely be identified with the L<sup>2</sup>-norm of the self-dual $`g`$-harmonic form denoted by $`c_1^+`$.
For each Riemannian metric $`g`$, the Seiberg-Witten equations
$`D_A\mathrm{\Phi }`$ $`=`$ $`0`$ (2)
$`F_A^+`$ $`=`$ $`i\sigma (\mathrm{\Phi })`$ (3)
are equations for an unknown Hermitian connection $`A`$ on $`L`$ and an unknown smooth section $`\mathrm{\Phi }`$ of $`๐_+`$. Here $`D_A`$ is the Dirac operator coupled to $`A`$, and $`\sigma :๐_+\mathrm{\Lambda }^+`$ is a certain canonical real-quadratic map. The latter formally arises from the fact that $`๐_+=๐_+L^{1/2}`$, whereas $`\mathrm{\Lambda }^+=^2๐_+`$; it is invariant under parallel transport, and satisfies $`|\sigma (\mathrm{\Phi })|^2=|\mathrm{\Phi }|^4/8`$. In conjunction with (3), the latter immediately gives us the important inequality
$$_M|\mathrm{\Phi }|^4๐\mu 32\pi ^2(c_1^+)^2$$
(4)
because $`2\pi c_1^+`$ is the harmonic part of $`\sigma (\mathrm{\Phi })`$.
In this paper, we will primarily be interested in $`4`$-manifolds for which there is a solution of the Seiberg-Witten equations for each metric. Let us make this more precise by introducing some terminology; cf. .
###### Definition 1
Let $`M`$ be a smooth compact oriented $`4`$-manifold, let $`HH^2(M,)`$ be a polarization of $`M`$, and let $`c`$ be a spin<sup>c</sup> structure on $`M`$. Then we will say that $`c`$ is a monopole class for $`(M,H)`$ if the Seiberg-Witten equations (23) have a solution for every $`H`$-adapted metric $`g`$.
This definition will be useful in practice, of course, only because of the existence of Seiberg-Witten invariants . For example, if a spin<sup>c</sup> structure satisfies $`[c_1(L)]^2=(2\chi +3\tau )(M)`$, and if $`c_1^+0`$ relative to the polarization $`H=_g^+`$, then the Seiberg-Witten invariant $`๐ฎ๐ฒ_c(M;H)`$ can be defined as the number of solutions, modulo gauge transformations and counted with orientations, of a generic perturbation
$`D_A\mathrm{\Phi }`$ $`=`$ $`0`$
$`iF_A^++\sigma (\mathrm{\Phi })`$ $`=`$ $`\varphi `$
of (23), where $`\varphi `$ is a smooth self-dual 2-form of small $`L^2`$ norm. If this invariant is non-zero, then $`c`$ is a monopole class for $`(M,H)`$, where $`H`$ is the polarization determined by the metric $`g`$. If $`b^+(M)2`$, $`๐ฎW_c(M,H)`$ is actually independent of the polarization $`H`$; when $`b^+(M)=1`$, by contrast, it is well defined only for those polarizations for which $`c_1^+0`$, and its value typically depends on whether $`c_1^+`$ is a future-pointing or past-pointing time-like vector in the Lorentzian vector space $`H^2(M,)`$.
More generally , suppose that we have a spin<sup>c</sup> structure such that
$$\mathrm{}=\frac{[c_1(L)]^2(2\chi +3\tau )(M)}{4}$$
is a non-negative integer. Then the moduli space $`_{c,g}`$ of gauge-equivalence classes of solutions of a generic perturbation of (23) is a smooth compact $`\mathrm{}`$-manifold; moreover, it acquires a canonical orientation once we choose an orientation for the vector space $`H^1(M,)H`$. Fix $`\mathrm{}`$ loops $`\beta _1,\mathrm{},\beta _{\mathrm{}}`$ in $`M`$, and define a smooth map $`_{c,g}T^{\mathrm{}}`$ from the moduli space to the $`\mathrm{}`$-torus by sending the gauge-equivalence class $`[(\mathrm{\Phi },A)]`$ to the holonomies of the $`U(1)`$ connection $`A`$ around the $`\mathrm{}`$ given loops. The homotopy class of this map only depends on the homology classes $`[\beta _i]H_1(M,)`$, one we may therefore define $`๐ฎ๐ฒ_c(M,[\beta _1],\mathrm{},[\beta _{\mathrm{}}];H)`$ to be the degree of this map. Again, if this invariant is non-zero, $`c`$ is a monopole class of $`(M,H)`$.
Many of the most remarkable consequences of Seiberg-Witten theory stem from the fact that the equations (23) imply the Weitzenbรถck formula
$$0=4^{}\mathrm{\Phi }+s\mathrm{\Phi }+|\mathrm{\Phi }|^2\mathrm{\Phi },$$
(5)
where $`s`$ denotes the scalar curvature of $`g`$. Taking the inner product with $`\mathrm{\Phi }`$, it follows that
$$0=2\mathrm{\Delta }|\mathrm{\Phi }|^2+4|\mathrm{\Phi }|^2+s|\mathrm{\Phi }|^2+|\mathrm{\Phi }|^4.$$
(6)
If we multiply (6) by $`|\mathrm{\Phi }|^2`$ and integrate, we have
$$0=_M\left[2\left|d|\mathrm{\Phi }|^2\right|^2+4|\mathrm{\Phi }|^2|\mathrm{\Phi }|^2+s|\mathrm{\Phi }|^4+|\mathrm{\Phi }|^6\right]๐\mu _g,$$
so that
$$(s)|\mathrm{\Phi }|^4๐\mu 4|\mathrm{\Phi }|^2|\mathrm{\Phi }|^2๐\mu +|\mathrm{\Phi }|^6๐\mu .$$
(7)
We will see in a moment that this implies some remarkable estimates for the Weyl curvature of suitable $`4`$-manifolds.
Before doing so, however, let us first recall that the self-dual Weyl tensor at a point $`x`$ of an oriented Riemannian $`4`$-manifold $`(M,g)`$ may be viewed as a trace-free endomorphism $`W_+(x):\mathrm{\Lambda }_x^+\mathrm{\Lambda }_x^+`$ of the self-dual $`2`$-forms at $`x`$. We will use $`w(x)`$ denotes its lowest eigenvalue; notice that this is automatically a Lipschitz continuous function $`w:M(\mathrm{},0]`$. Let us also fix the notational convention that, for any real-valued function $`f:M`$, $`f_{}:M(\mathrm{},0]`$ is defined by $`f_{}(x)=\mathrm{min}(f(x),0)`$. With these preliminaries, we are now prepared to prove our first, crucial result:
###### Proposition 2.1
Let $`(M,g)`$ be a compact oriented Riemannian $`4`$-manifold on which there is a solution of the Seiberg-Witten equations. Let $`c_1(L)`$ be the first Chern class of the relevant spin<sup>c</sup> structure, and let $`c_1^+`$ denote its self-dual part with respect to $`g`$. Then
$$V^{1/3}\left(_M|(\frac{2}{3}s_g+2w_g)_{}|^3๐\mu \right)^{2/3}32\pi ^2(c_1^+)^2,$$
(8)
where $`V=\text{Vol}(M,g)=_M๐\mu _g`$ is the total volume of $`(M,g)`$.
Proof. Any self-dual 2-form $`\psi `$ on any oriented 4-manifold satisfies the Weitzenbรถck formula
$$(d+d^{})^2\psi =^{}\psi 2W_+(\psi ,)+\frac{s}{3}\psi ,$$
where $`W_+`$ is the self-dual Weyl tensor. It follows that
$$_M(2W_+)(\psi ,\psi )_M(\frac{s}{3})|\psi |^2๐\mu _M|\psi |^2๐\mu ,$$
so that
$$_M2w|\psi |^2_M(\frac{s}{3})|\psi |^2๐\mu _M|\psi |^2๐\mu ,$$
and hence
$$_M(\frac{2}{3}s+2w)|\psi |^2_M(s)|\psi |^2๐\mu _M|\psi |^2๐\mu .$$
On the other hand, the particular self-dual 2-form $`\phi =\sigma (\mathrm{\Phi })=iF_A^+`$ satisfies
$`|\phi |^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}|\mathrm{\Phi }|^4,`$
$`|\phi |^2`$ $``$ $`{\displaystyle \frac{1}{2}}|\mathrm{\Phi }|^2|\mathrm{\Phi }|^2.`$
Setting $`\psi =\phi `$, we thus have
$$_M(\frac{2}{3}s+2w)|\mathrm{\Phi }|^4_M(s)|\mathrm{\Phi }|^4๐\mu 4_M|\mathrm{\Phi }|^2|\mathrm{\Phi }|^2๐\mu .$$
But (7) tells us that
$$_M(s)|\mathrm{\Phi }|^4๐\mu 4_M|\mathrm{\Phi }|^2|\mathrm{\Phi }|^2๐\mu _M|\mathrm{\Phi }|^6๐\mu ,$$
so we obtain
$$_M(\frac{2}{3}s+2w)_{}|\mathrm{\Phi }|^4๐\mu _M(\frac{2}{3}s+2w)|\mathrm{\Phi }|^4๐\mu _M|\mathrm{\Phi }|^6๐\mu .$$
By the Hรถlder inequality, we thus have
$$\left(|(\frac{2}{3}s+2w)_{}|^3๐\mu \right)^{1/3}\left(|\mathrm{\Phi }|^6๐\mu \right)^{2/3}|\mathrm{\Phi }|^6๐\mu ,$$
Since the Hรถlder inequality also tells us that
$$|\mathrm{\Phi }|^6๐\mu V^{1/2}\left(|\mathrm{\Phi }|^4๐\mu \right)^{3/2},$$
we thus have
$$V^{1/3}\left(_M|(\frac{2}{3}s_g+2w_g)_{}|^3๐\mu \right)^{2/3}|\mathrm{\Phi }|^4๐\mu 32\pi ^2(c_1^+)^2,$$
as claimed.
While this result is the key to everything that follows, its direct utility is limited by the fact that it is an $`L^3`$, rather than an $`L^2`$, estimate. Fortunately, however, we will be able to extract an $`L^2`$ estimate by means of a conformal rescaling trick, the general idea of which is drawn from Gursky :
###### Lemma 2.2
Let $`(M,\gamma )`$ be a compact oriented $`4`$-manifold with a fixed smooth conformal class of Riemannian metrics. Suppose, moreover, that $`\gamma `$ does not contain a metric of positive scalar curvature. Then, for any $`\alpha (0,1)`$, there is a metric $`g_\gamma \gamma `$ of differentiability class $`C^{2,\alpha }`$ for which $`s+3w`$ is a non-positive constant.
Proof. Let $`g_0\gamma `$ be a fixed smooth back-ground metric, and notice that, for $`g=u^2g_0\gamma `$, the function $`๐_g=s_g+3w_g`$ is given by
$$๐_g=u^3\left(6\mathrm{\Delta }_{g_0}u+๐_{g_0}u\right)$$
by virtue of the weighted conformal invariance of $`W_+`$. Thus $`๐`$ behaves under conformal rescaling just like the scalar curvature, despite the fact that it might well only be a Lipschitz function. In order to find a suitable choice of $`u`$, we therefore attempt to minimize
$$(u)=\frac{_M\left(6|du|^2+๐_{g_0}u^2\right)๐\mu _{g_0}}{\left(_Mu^4๐\mu _{g_0}\right)^{1/2}}$$
on the positive sector of the unit sphere in the Sobolev space $`L_1^2(M,g_0)`$. Yamabeโs ansatz for doing this is to minimize the functionals
$$_ฯต(u)=\frac{_M\left(6|du|^2+๐_{g_0}u^2\right)๐\mu _{g_0}}{\left(_Mu^{4ฯต}๐\mu _{g_0}\right)^{1/(2\frac{ฯต}{2})}}$$
and then take the limit of the minimizers as $`ฯต0`$. For each $`ฯต>0`$, the existence of minimizers follows from the Sobolev embedding theorem; indeed, following the proof of \[4, Theomem 5.5\], but deleting the very last sentence, one obtains a positive function $`u_ฯต`$ of class $`C^{2,\alpha }`$ which solves
$$6\mathrm{\Delta }_{g_0}u+๐_{g_0}u=c_ฯตu^{3ฯต},$$
(9)
where $`c_ฯต`$ is the infimum of $`_ฯต`$ on the positive sector of the unit sphere of $`L_1^2(M,g_0)`$. The convergence as $`ฯต0`$ then follows from an observation of Trudinger . Namely, since $`๐s`$ for any metric, and since there is nothing to prove if $`๐=s`$, our hypothesis that there is no metric of positive scalar curvature in $`\gamma `$ allows us to assume that that $`c_0=infc_ฯต<0`$. Inspection of (9) at a maximum then gives us the $`C^0`$ estimate
$$u_ฯต1+\sqrt{\left|\frac{inf๐_{g_0}}{c_0}\right|}$$
for all small $`\epsilon `$. This implies \[4, Theorem 6.5\] that $`u=lim_{ฯต0}u_ฯต`$ exists in $`C^1`$, and is a weak solution of
$$6\mathrm{\Delta }_{g_0}u+๐_{g_0}u=c_0u^3.$$
Schauder theory then tells us that $`u`$ is actually a $`C^{2,\alpha }`$ function. In particular, $`g_\gamma =u^2g_0`$ is a $`C^{2,\alpha }`$ metric, with curvature satisfying $`๐_{g_\gamma }=c_0`$ in the classical sense.
###### Theorem 2.3
Let $`(M,H)`$ be a polarized smooth compact oriented 4-manifold, and let $`c`$ be a monopole class for $`(M,H)`$. Let $`c_1(L)H^2(M,)`$ denote the anti-canonical class of $`c`$, and let $`c_1^+0`$ be its orthogonal projection to $`H`$ with respect to the intersection form. Then every $`H`$-adapted Riemannian metric $`g`$ satisfies
$$_M\left(\frac{2}{3}s2\sqrt{\frac{2}{3}}|W_+|\right)^2๐\mu 32\pi ^2(c_1^+)^2.$$
(10)
Proof. Let $`\gamma `$ be the conformal class of some $`H`$-adapted metric $`g`$. Since the Hodge star operator is conformally invariant in the middle dimension, every metric in $`\gamma `$ is also $`H`$-adapted. Since $`c`$ is a monopole class for $`H`$, it therefore follows that $`\gamma `$ does not contain any metrics of positive scalar curvature. Lemma 2.2 therefore applies, and tells us that the conformal class $`\gamma `$ contains a metric $`g_\gamma `$ for which that $`\frac{2}{3}s+2w=\frac{2}{3}๐`$ is a non-positive constant. For this metric, one then has
$$_M|(\frac{2}{3}s_{g_\gamma }+2w_{g_\gamma })_{}|^2๐\mu _{g_\gamma }=V_{g_\gamma }^{1/3}\left(_M|(\frac{2}{3}s_{g_\gamma }+2w_{g_\gamma })_{}|^3๐\mu _{g_\gamma }\right)^{2/3},$$
so that
$$_M|(\frac{2}{3}s_{g_\gamma }+2w_{g_\gamma })_{}|^2๐\mu _{g_\gamma }32\pi ^2(c_1^+)^2$$
by Proposition 2.1. Thus we at least have an $`L^2`$ estimate concerning the conformally related metric $`g_\gamma `$.
Let us now compare the left-hand side with analogous expression for the given metric $`g`$. To do so, we express $`g`$ in the form $`g=u^2g_\gamma `$, where $`u`$ is a positive $`C^2`$ function, and observe that
$`{\displaystyle _M}({\displaystyle \frac{2}{3}}s_g+2w_g)_{}u^2๐\mu _{g_\gamma }`$ $``$ $`{\displaystyle _M}\left({\displaystyle \frac{2}{3}}s_g+2w_g\right)u^2๐\mu _{g_\gamma }`$
$`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle ๐_gu^2๐\mu _{g_\gamma }}`$
$`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle u^3\left(6\mathrm{\Delta }_{g_\gamma }u+๐_{g_\gamma }u\right)u^2๐\mu _{g_\gamma }}`$
$`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle \left(6u^2|du|_{g_\gamma }^2+๐_{g_\gamma }\right)๐\mu _{g_\gamma }}`$
$``$ $`{\displaystyle \frac{2}{3}}{\displaystyle ๐_{g_\gamma }๐\mu _{g_\gamma }}`$
$`=`$ $`{\displaystyle _M}({\displaystyle \frac{2}{3}}s_{g_\gamma }+2w_{g_\gamma })_{}๐\mu _{g_\gamma }.`$
Applying Cauchy-Schwarz, we thus have
$`V_{g_\gamma }^{1/2}\left({\displaystyle |(\frac{2}{3}s_g+2w_g)_{}|^2๐\mu _g}\right)^{1/2}`$ $`=`$ $`V_{g_\gamma }^{1/2}\left({\displaystyle \left(\frac{2}{3}s_g+2w_g\right)^2u^4๐\mu _{g_\gamma }}\right)^{1/2}`$
$``$ $`{\displaystyle _M}\left({\displaystyle \frac{2}{3}}s_g+2w_g\right)u^2๐\mu _{g_\gamma }`$
$``$ $`{\displaystyle _M}\left({\displaystyle \frac{2}{3}}s_{g_\gamma }+2w_{g_\gamma }\right)๐\mu _{g_\gamma }`$
$`=`$ $`V_{g_\gamma }^{1/2}\left({\displaystyle |(\frac{2}{3}s_{g_\gamma }+2w_{g_\gamma })_{}|^2๐\mu _{g_\gamma }}\right)^{1/2},`$
and hence
$$_M|(\frac{2}{3}s_g+2w_g)_{}|^2๐\mu _g_M|(\frac{2}{3}s_{g_\gamma }+2w_{g_\gamma })_{}|^2๐\mu _{g_\gamma }.$$
This shows that
$$_M|(\frac{2}{3}s_g+2w_g)_{}|^2๐\mu _g32\pi ^2(c_1^+)^2$$
(11)
for every $`H`$-adapted metric $`g`$.
Finally, we observe that
$$\sqrt{\frac{2}{3}}|W_+|_gw_g$$
for any metric, simply because $`W_+`$ is trace-free. It follows that
$$\left(\frac{2}{3}s2\sqrt{\frac{2}{3}}|W_+|_g\right)^2|(\frac{2}{3}s_g+2w_g)_{}|^2$$
at every point, and (11) therefore implies the desired inequality (10).
The inequalities we have just derived are certainly sharp, in the sense that equality is attained by any Kรคhler metric of constant negative scalar curvature. However, we will now see that equality also holds in principle for a broader class of metrics. To make this precise, suppose that $`(M,\omega )`$ is a symplectic $`4`$-manifold, and let $`g`$ be a Riemannian metric such that
$$\omega (,)=g(J,)$$
for some almost complex structure $`J`$ on $`M`$; equivalently, suppose that $`\omega `$ is self-dual with respect to $`g`$ and has constant norm $`|\omega |_g=\sqrt{2}`$. One then says that $`g`$ is an almost-Kรคhler metric, with almost-Kรคhler form $`\omega `$. Evidently, $`g`$ is actually Kรคhler if $`J`$ is integrable, but Gromovโs theory of pseudo-holomorphic curves and Taubesโ characterization of the Seiberg-Witten invariants of symplectic $`4`$-manifolds have conclusively swept the study of such metrics into the mathematical main-stream.
In order to state our next result, we will need extra terminology in this connection. The $``$-scalar curvature of a $`4`$-dimensional almost-Kรคhler manifold $`(M,g,\omega )`$ is by definition the function
$$s^{}=\frac{s}{3}+2W_+(\omega ,\omega ).$$
This definition is motivated by the fact that the scalar curvature $`s`$ and $``$-scalar curvature $`s^{}`$ coincide for any Kรคhler manifold. Notice that we automatically have the inequality
$$s^{}\frac{s}{3}+4w,$$
with equality iff the almost-Kรคhler form $`\omega `$ belongs to the lowest eigenspace of $`W_+`$.
###### Proposition 2.4
Suppose that equality holds in either (8) or (11). Then $`g`$ is an almost-Kรคhler metric with the following properties:
* the almost-Kรคhler form $`\omega `$ belongs, at each $`xM`$, to the lowest eigenspace of $`W_+`$;
* the sum $`s+s^{}`$ of the scalar and $``$-scalar curvatures is a non-positive constant;
* the first Chern class of $`(M,\omega )`$ is of type $`(1,1)`$, and coincides with first Chern class $`c_1(L)`$ of the relevant spin<sup>c</sup> structure.
If equality holds in (10), these same conclusions hold, but, in addition,
* the self-dual Weyl curvature $`W_+`$ is invariant with respect to the almost-complex structure $`J`$ of $`(M,g,\omega )`$.
Conversely, any compact almost-Kรคhler $`4`$-manifold satisfying these conditions saturates the corresponding inequalities, and its first Chern class is a monopole class of the corresponding polarization.
Proof. If equality holds in (8), the last inequality in the proof of Proposition 2.1 forces $`\sigma (\mathrm{\Phi })`$ to be the harmonic representative of $`2\pi c_1^+`$. Moreover, the Hรถlder inequalities used in that proof shows that equality can only hold if $`|\mathrm{\Phi }|^2`$ and $`(\frac{2}{3}s+2w)`$ are constant, and moreover equal. Thus the harmonic form $`\phi =\sigma (\mathrm{\Phi })`$ has constant norm, so that $`\omega =\sqrt{2}\phi /|\phi |`$ is an almost-Kรคhler form compatible with $`g`$. Moreover, $`L=det(๐_+)=๐_+/\text{span}(\mathrm{\Phi })`$ coincides with the anti-canonical line bundle associated with the almost-complex structure $`J`$. Since $`c_1^+`$ is now a (non-positive) multiple of $`[\omega ]`$ and since the volume $`V=\text{Vol}(M,g)`$ of an almost-Kรคhler $`4`$-manifold is exactly $`[\omega ]^2/2`$, the fact that $`\frac{2}{3}s+2w`$ is a non-positive constant tells us that we now have
$$(\frac{2}{3}s+2w)๐\mu =V^{1/2}4\sqrt{2}\pi |c_1^+|=4\pi c_1(L)[\omega ],$$
since by assumption equality holds in (8). It thus follows that
$$\frac{1}{2}(s+s^{})๐\mu 4\pi c_1(L)[\omega ],$$
with equality iff $`\omega `$ everywhere belongs to the lowest eigenspace of $`W_+`$. However, Blair has shown that
$$\frac{1}{2}(s+s^{})๐\mu =4\pi c_1[\omega ]$$
(12)
for any almost-Kรคhler $`4`$-manifold. We thus conclude that $`\omega `$ is everywhere in the lowest eigenspace of $`W_+`$.
If we instead have equality in (11) or (10), we must in particular be at the minimum of
$$|(\frac{2}{3}s+2w)_{}|^2๐\mu $$
among metrics in the given conformal class. This then implies that $`\frac{2}{3}s+2w`$ is constant, and we therefore also have equality in (8), and the claim follows from the previous argument.
The converse assertions follow from and (12). Details are left to the interested reader.
In a naรฏve sense, this provides a complete characterization of the metrics which saturate (8), (10), and (11). It remains to be seen, however, whether such metrics can ever be strictly almost-Kรคhler, in the sense that the almost-complex structure $`J`$ fails to be integrable. For example, a beautiful recent result of J. Armstrong asserts the non-existence of compact, Einstein, strictly almost-Kรคhler $`4`$-manifolds on which $`\omega `$ is everywhere an eigenvector of $`W_+`$. Thus:
###### Corollary 2.5
Suppose that $`g`$ is an Einstein metric which saturates (8), (10), or (11). Then $`g`$ is actually Kรคhler-Einstein.
Similarly, a recent result of Apostolov-Armstrong-Drฤghici implies that an almost-Kรคhler metric saturating (10) is Kรคhler iff its Ricci tensor is $`J`$-invariant.
## 3 Einstein Metrics
Our first application of the preceding Weyl estimates will be to prove new non-existence for Einstein metrics on suitable smooth compact $`4`$-manifolds. We begin by proving the following technical result:
###### Proposition 3.1
Let $`(M,H)`$ be a polarized smooth compact oriented 4-manifold with monopole class $`c`$. Then every $`H`$-adapted metric satisfies
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+2|W_+|_g^2\right)๐\mu _g\frac{2}{3}(c_1^+)^2.$$
If equality holds, moreover, $`g`$ is an almost-Kรคhler metric of the type described in Proposition 2.4.
Proof. Let us begin by rewriting the inequality (10) as
$$\frac{2}{3}s2\sqrt{\frac{2}{3}}|W_+|4\sqrt{2}\pi |c_1^+|,$$
where $``$ denotes the $`L^2`$ norm with respect to $`g`$. By the triangle inequality, we therefore have
$$\frac{2}{3}s+\frac{1}{3}(\sqrt{24}|W_+|)4\sqrt{2}\pi |c_1^+|.$$
(13)
We now elect to interpret the left-hand side as the dot product
$$(\frac{2}{3},\frac{1}{3\sqrt{2}})(s,(\sqrt{48}|W_+|))=\frac{2}{3}s+\frac{1}{3}(\sqrt{24}|W_+|)$$
in $`^2`$. Applying Cauchy-Schwarz, we thus have
$$\left((\frac{2}{3})^2+(\frac{1}{3\sqrt{2}})^2\right)^{1/2}\left(_M(s^2+48|W_+|^2)๐\mu \right)^{1/2}\frac{2}{3}s+\frac{1}{3}(\sqrt{24}|W_+|).$$
Thus
$$\frac{1}{2}_M(s^2+48|W_+|^2)๐\mu 32\pi ^2(c_1^+)^2,$$
so that
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+2|W_+|_g^2\right)๐\mu _g\frac{2}{3}(c_1^+)^2,$$
as claimed. If equality holds, moreover, (10) is saturated, and Proposition 2.4 applies.
###### Proposition 3.2
Let $`X`$ be a compact oriented $`4`$-manifold with a non-trivial Seiberg-Witten invariant, and set
$$M=X\mathrm{\#}k\overline{}_2\mathrm{\#}\mathrm{}(S^1\times S^3)$$
for integers $`k,\mathrm{}0`$. Then any Riemannian metric $`g`$ on $`M`$ satisfies
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+2|W_+|_g^2\right)๐\mu _g\frac{2}{3}(2\chi +3\tau )(X),$$
and equality is possible only if $`k=\mathrm{}=0`$.
Proof. The proof is a direct extension of the computations in , although for $`\mathrm{}>0`$ we now use the holonomy-constrained Seiberg-Witten invariant of ; cf. .
Now notice that we may assume that $`(2\chi +3\tau )(X)>0`$, since otherwise there is nothing to prove. This has the pleasant consequence that any Seiberg-Witten invariant of $`X`$ is independent of polarization, even if $`b^+(X)=1`$. Let $`c_1(X)`$ denote the first Chern class of a spin<sup>c</sup> structure on $`X`$ for which the Seiberg-Witten invariant is non-zero, and notice that $`(c_1(X))^2(2\chi +3\tau )(X)>0`$, because the relevant Seiberg-Witten moduli space must have non-negative virtual dimension. Pull $`c_1(X)`$ back to $`M=X\mathrm{\#}k\overline{}_2\mathrm{}(S^1\times S^3)`$ via the canonical collapsing map, and, by a standard abuse of notation, let $`c_1(X)`$ also denote this pulled-back class. Thus, with respect to our given polarization,
$$([c_1(X)]^+)^2(c_1(X))^2(2\chi +3\tau )(X)>0.$$
Now choose generators $`E_1,\mathrm{},E_k`$ for the pull-backs to $`M`$ of the $`k`$ relevant copies of $`H^2(\overline{}_2,)`$ so that
$$[c_1(X)]^+E_j0,j=1,\mathrm{},k.$$
Let $`\beta _1,\mathrm{},\beta _{\mathrm{}}`$ be closed curves in $`M`$ which generate of the fundamental groups of the $`\mathrm{}`$ relevant copies of $`S^1\times S^3`$. Then there is a spin<sup>c</sup> structure on $`M`$ with $`๐ฎ๐ฒ_c(M,[\beta _1],\mathrm{},[\beta _{\mathrm{}}];H)0`$ and
$$c_1(L)=c_1(X)\underset{j=1}{\overset{k}{}}E_j.$$
Thus $`c`$ is a monopole class of $`(M,H)`$. But one then has
$`(c_1^+)^2`$ $`=`$ $`\left([c_1(X)]^++{\displaystyle \underset{j=1}{\overset{k}{}}}E_j^+\right)^2`$
$`=`$ $`([c_1(X)]^+)^22{\displaystyle \underset{j=1}{\overset{k}{}}}[c_1(X)]^+E_j+({\displaystyle \underset{j=1}{\overset{k}{}}}E_j^+)^2`$
$``$ $`([c_1(X)]^+)^2`$
$``$ $`(2\chi +3\tau )(X),`$
exactly as claimed.
If equality held, $`g`$ would be almost-Kรคhler, and $`c_1(L)`$ would be the anti-canonical class of the associated almost-complex structure on $`M`$. But, by construction,
$$c_1^2(L)=(c_1(X)E_j)^2=(2\chi +3\tau )(X)k,$$
whereas
$$(2\chi +3\tau )(M)=(2\chi +3\tau )(X)k4\mathrm{}$$
by Mayer-Vietoris. Since $`c_1^2=2\chi +3\tau `$ for an almost-complex manifold, equality is thus excluded unless $`\mathrm{}=0`$.
There is yet more information available, however. Indeed, if equality held, the almost-Kรคhler class $`[\omega ]`$ would also necessarily be a non-positive multiple of $`c_1^+`$. On the other hand, our computation shows that equality can only hold if $`[c_1(X)]^+E_j=0`$, so it would follow that $`[\omega ]E_j=0`$ for all $`j`$. However, the Seiberg-Witten invariant would then also be non-trivial for a spin<sup>c</sup> structure with $`c_1(\stackrel{~}{L})=c_1(L)+2E_1`$, and a celebrated theorem of Taubes would then force the homology class $`E_j`$ to be represented by a pseudo-holomorphic $`2`$-sphere in the symplectic manifold $`(M,\omega )`$. But the (positive!) area of this sphere with respect to $`g`$ would then be exactly $`[\omega ]E_j`$, contradicting the observation that $`[\omega ]E_j=0`$. Thus equality can definitely be excluded unless $`k`$ and $`\mathrm{}`$ both vanish.
This then gives us a new non-existence theorem for Einstein metrics:
###### Theorem 3.3
Let $`X`$ be a compact oriented $`4`$-manifold with a non-trivial Seiberg-Witten invariant. Then
$$M=X\mathrm{\#}k\overline{}_2\mathrm{\#}\mathrm{}(S^1\times S^3)$$
does not admit Einstein metrics if $`k+4\mathrm{}\frac{1}{3}(2\chi +3\tau )(X)`$.
Proof. We may assume that $`(2\chi +3\tau )(X)>0`$, since otherwise the result follows from the Hitchin-Thorpe inequality.
Now
$$(2\chi +3\tau )(M)=\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+2|W_+|_g^2\frac{|\stackrel{}{r}|^2}{2}\right)๐\mu _g$$
for any metric on $`g`$ on $`M`$. If $`g`$ is an Einstein metric, the trace-free part $`\stackrel{}{r}`$ of the Ricci curvature vanishes, and we then have
$`(2\chi +3\tau )(X)k4\mathrm{}`$ $`=`$ $`(2\chi +3\tau )(M)`$
$`=`$ $`{\displaystyle \frac{1}{4\pi ^2}}{\displaystyle _M}\left({\displaystyle \frac{s_g^2}{24}}+2|W_+|_g^2\right)๐\mu _g`$
$``$ $`{\displaystyle \frac{2}{3}}(2\chi +3\tau )(X)`$
by Proposition 3.2, with equality only if $`k`$ and $`\mathrm{}`$ both vanish. If $`M`$ carries an Einstein metric, it therefore follows that
$$\frac{1}{3}(2\chi +3\tau )(X)>k+4\mathrm{}.$$
The claim now follows by contraposition.
Specializing to the case of complex surfaces, we now have:
###### Corollary 3.4
Let $`(X,J_X)`$ be a minimal complex surface of general type, and let $`(M,J_M)`$ be obtained from $`X`$ by blowing up $`k`$ points. Then the smooth compact $`4`$-manifold $`M`$ does not admit any Einstein metrics if $`k\frac{1}{3}c_1^2(X)`$.
Example Let $`X_3`$ be a hypersurface of degree $`6`$. Since the canonical class on $`X`$ is twice the hyperplane class, $`c_1^2(X)=2^26=24`$. Corollary 3.4 therefore tells us that if we blow up $`X`$ at $`8`$ points, the resulting $`4`$-manifold
$$M=X\mathrm{\#}8\overline{}_2$$
does not admit Einstein metrics.
Now let us compare the complex surface $`M`$ with the Horikawa surface $`N`$ obtained as a ramified double cover of $`_1\times _1`$ branched at a generic curve of bidegree $`(6,12)`$. Both of these simply connected complex surfaces have $`c_1^2=16`$ and $`p_g=10`$, and the underlying oriented $`4`$-manifolds therefore have $`b_+=21`$ and $`b_{}=93`$. In particular, both have signature $`\tau =720mod16`$, so, by Rochlinโs theorem, neither is spin. Thus $`M`$ and $`N`$ have isomorphic intersection forms by the Minkowski-Hasse classification, and are therefore homeomorphic by Freedmanโs theorem . However, $`N`$ has ample canonical line bundle, and so admits a Kรคhler-Einstein metric by Yauโs theorem. Thus, although $`M`$ and $`N`$ are homeomorphic, one admits Einstein metrics, while the other doesnโt.
If we instead start with a hypersurface $`X_3`$ of degree $`7`$, one can construct homeomorphic pairs $`(M,N)`$ for which the minimal complex surface sits well above the Noether line. Details are left to the interested reader. $`\mathrm{}`$
Remark Corollary 3.4 is the direct descendant of an analogous result in , where, using only scalar curvature estimates, a similar conclusion was proved for $`k\frac{2}{3}c_1^2(X)`$. It was later pointed out by Kotschick that such a result alone suffices to imply the existence of homeomorphic pairs consisting of an Einstein manifold and a $`4`$-manifold which does not admit Einstein metrics; however, the examples that arise by this method are quite complicated, and have huge $`c_1^2`$.
The intermediate step between and Corollary 3.4 may be found in , where Seiberg-Witten estimates of Weyl curvature were first introduced. While crude by present standards, the method used there did lead to an obstruction when $`k\frac{25}{57}c_1^2(X)`$, or about two-thirds of the way to the present result.
The fact that $`S^1\times S^3`$ handles can be added to a $`4`$-manifold without losing Seiberg-Witten control of the scalar curvature was first observed by Petean , although Seiberg-Witten theory only enters his result in an indirect manner. The search for a direct Seiberg-Witten explanation of this phenomenon then led the author and del Rio to a discovery of the generalized Seiberg-Witten invariant used here. In fact, however, this turned out to merely be a re-discovery; quite different considerations had already led Ozsvรกth and Szabรณ to develop a full-blown theory, described in a preprint written the previous month. $`\mathrm{}`$
Let us now recall that the Hitchin-Thorpe inequality asserts that if a compact oriented $`4`$-manifold $`M`$ admits an Einstein metric, then
$$\frac{|\tau (M)|}{\chi (M)}\frac{2}{3}.$$
Hitchin went on to observe that the converse is certainly false, because the argument shows in particular that a simply connected $`4`$-manifold with $`2\chi (M)=3\tau (M)`$ can admit an Einstein metric only if its a $`K3`$ surface. Examples of $`4`$-manifolds with $`|\tau (M)|/\chi (M)<2/3`$ which do not admit Einstein metrics were first constructed by Gromov , using his simplicial volume invariant; however, this method only works in the presence of an infinite fundamental group. In , simply connected examples were first constructed, using Seiberg-Witten estimates for the scalar curvature. Shortly thereafter, Sambusetti showed that the entropy estimates of Besson-Courtois-Gallot allow one to construct examples of arbitrary Euler characteristic and signature, but again with huge fundamental group; see Petean and del Rio for similar results. In the simply connected case, however, it remains to be seen whether $`(\chi ,\tau )`$ can really be arbitrarily specified, subject to the obvious constraints $`\chi 2+|\tau |`$ and $`\chi \tau mod2`$. Indeed, even the following is open:
###### Question 3.5
For every $`q(1,1)`$, are there smooth, compact simply connected $`4`$-manifolds with $`\tau /\chi =q`$ which do not admit Einstein metrics?
Hitchin-Thorpe gives one a resoundingly affirmative answer in the range $`\frac{2}{3}|q|<1`$. The present results allow us to improve this as follows:
###### Corollary 3.6
Let $`q`$ be a rational number with $`\frac{8}{23}|q|<1`$. Then there are smooth, compact, simply connected $`4`$-manifolds with $`\tau /\chi =q`$ which do not admit Einstein metrics.
Proof. For $`m`$ any even integer bigger than $`17`$ million, Chen has constructed a simply connected minimal complex surface $`X`$ of general type (in fact, a hyperelliptic fibration) with $`\tau (X)=m`$ and $`\chi (X)=4m`$. If we now blow up such a surface at $`k`$ points, where $`k\frac{11}{3}m`$, the resulting simply connected $`4`$-manifold $`M=X\mathrm{\#}k\overline{}_2`$ does not admit Einstein metrics by Corollary 3.4. Now for this manifold we have
$$\frac{\tau (M)}{\chi (M)}=\frac{5}{4+\frac{k}{m}}1,$$
and the right-hand side sweeps out $`(1,\frac{8}{23}]`$ as $`\frac{k}{m}`$ ranges over $`[\frac{11}{3},\mathrm{})`$. This proves the claim for $`q`$ negative. The case of $`q`$ positive then follows by reversing orientation.
The Weyl estimates of ยง2 also have interesting ramifications for the theory of anti-self-dual $`4`$-manifolds. Recall that an oriented Riemannian $`4`$-manifold is said to be anti-self-dual if it satisfies the $`W_+0`$, and that this condition is conformally invariant. Compact anti-self-dual manifolds exist in profusion. Indeed, Taubes has shown that for any smooth compact orientable $`X^4`$, there is an integer $`k_0`$ such that $`M=X\mathrm{\#}k\overline{}_2`$ admits metrics with $`W_+=0`$ provided that $`kk_0`$. In particular, if we blow up a symplectic $`4`$-manifold with $`b^+>1`$ at enough points, we obtain a manifold which admits anti-self-dual metrics and also has a non-trivial Seiberg-Witten invariant. Let us now consider the scalar curvature of such manifolds.
###### Lemma 3.7
Let $`(M,g)`$ be a compact anti-self-dual 4-manifold with a non-zero Seiberg-Witten invariant. Then
$$_Ms^2๐\mu 72\pi ^2(c_1^+)^2,$$
where $`c_1^+`$ is again the self-dual part of the first Chern class of the relevant spin<sup>c</sup> structure. If equality holds, moreover, $`g`$ is an almost-Kรคhler metric of the type describe in Proposition 2.4.
Proof. If $`W_+0`$, (10) becomes
$$_M\frac{4}{9}s^2๐\mu 32\pi ^2(c_1^+)^2,$$
and the desired inequality is therefore an immediate consequence. Moreover, if equality holds, Proposition 2.4 applies, and $`g`$ is an almost-Kรคhler manifold with the relevant special properties.
Notice that this result may be used as a vanishing theorem. For example, if $`c`$ is a spin<sup>c</sup> structure on a hyperbolic $`4`$-manifold, the corresponding Seiberg-Witten invariant will vanish unless $`(c_1^+)^2\frac{8}{3}\chi `$. Since the non-negativity of the virtual dimension of the Seiberg-Witten moduli space would require that $`(c_1^+)^22\chi +|c_1^{}|^2`$, this would seem to lend some weak support to the conjecture that the Seiberg-Witten invariants of a hyperbolic $`4`$-manifold are all zero. In any case, it also implies the following result:
###### Theorem 3.8
Let $`X`$ be a compact oriented $`4`$-manifold with a non-trivial Seiberg-Witten invariant. Then
$$M=X\mathrm{\#}k\overline{}_2\mathrm{\#}\mathrm{}(S^1\times S^3)$$
does not admit anti-self-dual Einstein metrics if $`k+4\mathrm{}\frac{1}{4}(2\chi +3\tau )(X)`$.
Proof. As in the proof of Proposition 3.2, there is a monopole class $`c`$ of $`(M,H)`$ with
$$c_1(L)=c_1(X)\underset{j=1}{\overset{k}{}}E_j,$$
and such that
$$(c_1^+)^2(2\chi +3\tau )(X).$$
If $`M`$ carried an anti-self-dual Einstein metric, Lemma 3.7 would then tell us that
$`(2\chi +3\tau )(X)k4\mathrm{}`$ $`=`$ $`(2\chi +3\tau )(M)`$
$`=`$ $`{\displaystyle \frac{1}{4\pi ^2}}{\displaystyle _M}{\displaystyle \frac{s^2}{24}}๐\mu `$
$``$ $`{\displaystyle \frac{72\pi ^2}{96\pi ^2}}(c_1^+)^2`$
$``$ $`{\displaystyle \frac{3}{4}}(2\chi +3\tau )(X),`$
so that
$$k+4\mathrm{}\frac{1}{4}(2\chi +3\tau )(X).$$
Moreover, equality could only hold if $`k`$ and $`\mathrm{}`$ both vanished for precisely the same reasons delineated in the proof of Theorem 3.3.
The result now follows by contraposition.
## 4 Ricci Curvature Estimates
One of the most striking aspects of Seiberg-Witten theory is that it allows one to calculate the precise value of the invariant
$$_s(M)=\underset{g}{inf}_Ms_g^2๐\mu _g$$
for large classes of $`4`$-manifolds $`M`$; for example, if $`M`$ is a complex surface of general type, and if $`X`$ is its minimal model, then
$$\underset{g}{inf}_Ms_g^2๐\mu _g=32\pi ^2c_1^2(X).$$
(14)
However, one would also like to compute the corresponding infimum
$$_r(M)=\underset{g}{inf}_M|r_g|^2๐\mu _g$$
(15)
for the Ricci curvature $`r`$. Since
$$|r|^2d\mu =(\frac{s^2}{4}+|\stackrel{}{r}|^2)d\mu ,$$
where $`\stackrel{}{r}`$ is the trace-free part of the Ricci curvature, we obviously have
$$_r(M)\frac{1}{4}_s(M)$$
(16)
and the scalar curvature estimate obviously gives us some important information about this problem. Moreover, because Kรคhler-Einstein metrics saturate the lower bound for the scalar curvature, one actually gets equality in (16) if $`M`$ happens to be a minimal complex surface of general type. For non-minimal surfaces, however, the situation is dramatically different; indeed, even the relatively crude Weyl estimates of allow one to deduce that, for any complex surface $`M`$ of general type,
equality holds in (16) $``$ $`M`$ is minimal.
Until now, however, non-trivial direct calculations of $`_r`$ have remained beyond the scope of the theory. Fortunately, we will now see that the estimates of ยง2 are perfectly suited to this purpose. The key observation is the following:
###### Proposition 4.1
Let $`(M,H)`$ be a polarized smooth compact oriented 4-manifold with monopole class $`c`$. Then every $`H`$-adapted metric satisfies
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g\frac{1}{2}(c_1^+)^2.$$
Moreover, equality holds iff $`g`$ is a Kรคhler metric of constant scalar curvature, and the spin<sup>c</sup> structure is the one determined by a $`g`$-compatible complex structure on $`M`$.
Proof. We again begin with inequality (13),
$$\frac{2}{3}s+\frac{1}{3}(\sqrt{24}|W_+|)4\sqrt{2}\pi |c_1^+|,$$
but this time interpret the left-hand side as the dot product
$$(\frac{2}{3},\frac{\sqrt{2}}{3})(s,(\sqrt{12}|W_+|))$$
in $`^2`$. Applying the Cauchy-Schwarz inequality, we then obtain
$$\left((\frac{2}{3})^2+(\frac{\sqrt{2}}{3})^2\right)^{1/2}\left(_M(s^2+12|W_+|^2)๐\mu \right)^{1/2}\frac{2}{3}s+\frac{1}{3}(\sqrt{24}|W_+|).$$
Thus
$$\frac{2}{3}_M(s^2+12|W_+|^2)๐\mu 32\pi ^2(c_1^+)^2,$$
and
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g\frac{1}{2}(c_1^+)^2,$$
as claimed.
If equality holds, our use of Cauchy-Schwarz forces
$$(\frac{2}{3}:\frac{\sqrt{2}}{3})=(s:(\sqrt{12}|W_+|)),$$
which is to say that
$$|W_+|^2๐\mu =\frac{s^2}{24}๐\mu .$$
In this case, we then have
$$_Ms_g^2๐\mu _g=32\pi ^2(c_1^+)^2,$$
and it then follows that the metric is Kรคhler, and has constant scalar curvature. Conversely, a Kรคhler surface with constant scalar curvature satisfies $`s\sqrt{24}|W_+|`$ and $`s^2๐\mu =32\pi ^2(c_1^+)^2`$, so that equality is attained for any such metric.
This then gives rise to a Ricci curvature estimate via Gauss-Bonnet formulรฆ :
###### Proposition 4.2
Let $`(M,H)`$ be a polarized smooth compact oriented 4-manifold with monopole class $`c`$. Then every $`H`$-adapted metric satisfies
$$\frac{1}{8\pi ^2}_M|r_g|^2๐\mu _g2(c_1^+)^2(2\chi +3\tau )(M),$$
where $`r`$ denotes the Ricci curvature of $`g`$.
Proof. Since (1) tells us that
$$(2\chi +3\tau )(M)=\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+2|W_+|_g^2\frac{|\stackrel{}{r}|^2}{2}\right)๐\mu _g,$$
it follows that
$$\frac{1}{8\pi ^2}_M|r|^2๐\mu =\frac{1}{\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g(2\chi +3\tau )(M)$$
(17)
for any metric on $`g`$ on $`M`$. The result now follows immediately from Proposition 4.1.
This observation supplies the key to the main result of this section:
###### Theorem 4.3
Let $`X`$ be any minimal complex surface of general type, and let $`M=X\mathrm{\#}k\overline{}_2\mathrm{\#}\mathrm{}(S^1\times S^3)`$. Then
$$_r(M)=\underset{g}{inf}_M|r|^2๐\mu =8\pi ^2(c_1^2(X)+k+4\mathrm{}).$$
Proof. Since
$$(2\chi +3\tau )(M)=c_1^2(X)k4\mathrm{},$$
equation (17) tells us that it suffices to prove that
$$\underset{g}{inf}\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g=\frac{1}{2}c_1^2(X).$$
Now observe that the proof of Proposition 3.2 tells us that that, for every metric $`g`$ on $`M`$, there is a spin<sup>c</sup> structure $`c`$ with non-zero Seiberg-Witten invariant such that
$$(c_1^+)^2(2\chi +3\tau )(X)=c_1^2(X).$$
Hence
$$\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g\frac{1}{2}c_1^2(X)$$
for every metric $`g`$ on $`M`$. It follows that
$$\underset{g}{inf}\frac{1}{4\pi ^2}_M\left(\frac{s_g^2}{24}+\frac{1}{2}|W_+|_g^2\right)๐\mu _g\frac{1}{2}c_1^2(X).$$
To finish the proof, it suffices to produce a family of metrics $`g_\epsilon `$ on $`M`$ such that
$$\underset{\epsilon 0^+}{lim\; sup}\frac{1}{4\pi ^2}_M\left(\frac{s_{g_\epsilon }^2}{24}+\frac{1}{2}|W_+|_{g_\epsilon }^2\right)๐\mu _{g_\epsilon }\frac{1}{2}c_1^2(X).$$
To do this, let $`\stackrel{ห}{X}`$ be the pluri-canonical model of $`X`$, which carries a Kรคhler-Einstein orbifold metric $`\stackrel{ห}{g}`$ by an immediate generalization of the Aubin/Yau solution of the $`c_1<0`$ case of the Calabi conjecture. Now
$$_{\stackrel{ห}{X}}\left(\frac{s_{\stackrel{ห}{g}}^2}{24}+\frac{1}{2}|W_+|_{\stackrel{ห}{g}}^2\right)๐\mu _{\stackrel{ห}{g}}=\frac{1}{2}c_1^2(\stackrel{ห}{X})=\frac{1}{2}c_1^2(X)$$
because $`|W_+|^2s^2/24`$ for any Kรคhler surface, whereas a Kรคhler-Einstein surface has Ricci form given by $`s\omega /4`$, where $`\omega `$ is the Kรคhler form. Let $`o_1,\mathrm{},o_n`$ be the orbifold points of $`\stackrel{ห}{X}`$, if there are any. Each of these singularities is of type A-D-E, meaning that $`o_j`$ has a neighborhood is modeled on $`^2/\mathrm{\Gamma }_j`$ for some discrete sub-group $`\mathrm{\Gamma }_j`$ of $`SU(2)`$. The passage from $`\stackrel{ห}{X}`$ to $`X`$ is accomplished by replacing each $`o_j`$ by a collection of $`(2)`$-curves with intersections determined by the Dynkin diagram (Coxeter graph) of type A-D-E corresponding to $`\mathrm{\Gamma }_j`$.
One can now construct smooth metrics on the minimal model $`X`$ by modifying the orbifold metric $`\stackrel{ห}{g}`$, without introducing substantial amounts of extra volume, Ricci curvature, or self-dual Weyl curvature. Indeed, for each $`\mathrm{\Gamma }_j`$ there are gravitational instanton metrics on the minimal resolution of $`^2/\mathrm{\Gamma }_j`$, which is precisely obtained by replacing the origin with a set of $`(2)`$-curves as above. These gravitational instanton metrics are Ricci-flat and anti-self-dual, and they closely approximate the Euclidean metric on $`^2/\mathrm{\Gamma }_j`$, being of the form $`\delta _{jk}+O(\varrho ^4)`$ in the asymptotic region, where $`\varrho `$ is the Euclidean radius; moreover, the first and second derivatives of such a metric are of order $`O(\varrho ^5)`$ and $`O(\varrho ^6)`$, respectively, in these coordinates. Fix such a metric, and consider the family of gravitational instanton metrics $`h_\epsilon `$ obtained by multiplying it by $`\epsilon ^4`$, and then making the homothetic change $`x\epsilon ^2x`$ of asymptotic coordinates. The metric $`h_\epsilon `$ is then uniformly $`\delta _{jk}+O(\epsilon ^8\varrho ^4)`$ on the complement of the ball of radius $`\epsilon ^2`$, with first and second derivatives $`O(\epsilon ^8\varrho ^5)`$ and $`O(\epsilon ^8\varrho ^6)`$, respectively. Take geodesic spray coordinates about each of the orbifold points $`o_j`$, so that $`\stackrel{ห}{g}`$ takes the form $`\delta _{jk}+O(\varrho ^2)`$ in these coordinates. Delete the $`\epsilon `$ ball around each orbifold point, and replace it in the $`\varrho \epsilon `$ region of the corresponding gravitational instanton; and on the transitional annulus $`\varrho [\epsilon ,2\epsilon ]`$, consider the metric
$$g_\epsilon =f(\frac{\varrho }{\epsilon })h_\epsilon +[1f(\frac{\varrho }{\epsilon })]\stackrel{ห}{g},$$
(18)
where $`f:[1,2][0,1]`$ is a fixed smooth function which is $`1`$ near $`1`$ and $`0`$ near $`2`$. This gives us a family of metrics on $`X`$, the geometry of which will be analyzed in a moment.
However, what we really need is a metric on $`M`$ rather than on $`X`$, so we need to modify $`\stackrel{ห}{g}`$ a bit more. Now $`S^1\times S^3`$ and $`\overline{}_2`$ both admit anti-self-dual metrics of positive scalar curvature โ namely, the obvious product metric on $`S^1\times S^3`$, and the Fubini-Study metric on the reverse-oriented complex projective plane $`\overline{}_2`$. The complement of any point in $`S^1\times S^3`$ or $`\overline{}_2`$ therefore admits an asymptotically flat metric with $`s0`$ and $`W_+0`$, namely the standard metric rescaled by the Yamabe Greens function of the appropriate point. One can then find asymptotic charts for these manifolds, parameterized by the complement of, say, the unit ball in $`^4`$, in which these metrics differ from the Euclidean metric by $`O(\varrho ^2)`$ and in which first and second partial derivatives are respectively $`O(\varrho ^3)`$ and $`O(\varrho ^4)`$. (Notice the fall-off rate is slower than in the previous case; these metrics have positive mass!) Again multiplying by $`\epsilon ^4`$, where $`\epsilon `$ is to be viewed as an auxiliary parameter, and making a homothety as before, we thus get metrics on the complement of the $`\epsilon ^2`$ ball in $`^4`$ of the form $`\delta _{jk}+O(\epsilon ^4\varrho ^2)`$, with first and second partial derivatives of order $`O(\epsilon ^4\varrho ^3)`$ and $`O(\epsilon ^4\varrho ^4)`$, respectively. Again, let $`h_\epsilon `$ denote the restriction of the resulting metric on the the $`\varrho >2\epsilon `$ regions of $`k`$ disjoint punctured copies of $`\overline{}_2`$ and $`\mathrm{}`$ disjoint punctured copies of $`S^1\times S^3`$. Take geodesic spray coordinates around non-singular points $`p_1,\mathrm{},p_k,q_1,\mathrm{},q_{\mathrm{}}`$ of $`\stackrel{ห}{X}`$, and delete $`\epsilon `$-balls around each one, and define a metric $`g_\epsilon `$ on $`M`$ as $`h_\epsilon `$ when $`\varrho <\epsilon `$, as $`\stackrel{ห}{g}`$ when $`\varrho >2\epsilon `$, and as (18) on the transitional annuli $`\varrho [\epsilon ,2\epsilon ]`$. The sectional curvature of the metrics $`g_\epsilon `$ are then uniformly bounded on the annuli $`\epsilon \varrho 2\epsilon `$ as $`\epsilon 0`$, while the volumes of these annuli are $`O(\epsilon ^4)`$ as $`\epsilon 0`$. Since the metrics we have used to replace the $`\epsilon `$-balls are all scalar-flat and anti-self-dual, we therefore have
$$\frac{1}{4\pi ^2}_M\left(\frac{s_{g_\epsilon }^2}{24}+\frac{1}{2}|W_+|_{g_\epsilon }^2\right)๐\mu _{g_\epsilon }C\epsilon ^4+\frac{1}{4\pi ^2}_{\stackrel{ห}{X}}\left(\frac{s_{\stackrel{ห}{g}}^2}{24}+\frac{1}{2}|W_+|_{\stackrel{ห}{g}}^2\right)๐\mu _{\stackrel{ห}{g}},$$
and it thus follows that
$$\underset{\epsilon 0^+}{lim\; sup}\frac{1}{4\pi ^2}_M\left(\frac{s_{g_\epsilon }^2}{24}+\frac{1}{2}|W_+|_{g_\epsilon }^2\right)๐\mu _{g_\epsilon }\frac{1}{2}c_1^2(X).$$
The theorem follows.
###### Corollary 4.4
Let $`M`$ be the underlying $`4`$-manifold of any compact complex surface of Kodaira dimension $`0`$. Let $`X`$ be the minimal model of $`M`$, and let $`k`$ be the number of points at which $`X`$ must be blown up so as to obtain $`M`$. Then
$$\underset{g}{inf}_M|r|^2๐\mu =8\pi ^2(c_1^2(X)+k),$$
where the infimum is taken over all smooth Riemannian metrics on $`g`$ on $`M`$.
Proof. Complex surfaces of Kodaira dimension $`2`$, are, by definition, precisely those of general type; for these, the result is just the $`\mathrm{}=0`$ case of Theorem 4.3. On the other hand, the analogous assertion for Kodaira dimensions $`0`$ and $`1`$ immediately follows from the fact that any elliptic surface admits sequences of metrics for which $`\text{Vol}(M,g)0`$, but for which $`s`$ and $`|W_+|`$ are uniformly bounded.
In particular, in view of (14), this gives a satisfying new proof of the following:
###### Corollary 4.5
Let $`M`$ be the underlying $`4`$-manifold of any complex surface of Kodaira dimension $`0`$. Then
equality holds in (16) $``$ $`M`$ is minimal.
We are also see that the invariant $`_r`$ is extremely sensitive to changes in differentiable structure on a topological $`4`$-manifold:
###### Corollary 4.6
Let $`X`$ and $`\stackrel{~}{X}`$ be two minimal simply connected complex surfaces with the same geometric genus $`p_g0`$, but with $`c_1^2(X)c_1^2(\stackrel{~}{X})=j>0`$. Then for all positive integers $`k`$, the $`4`$-manifolds $`\stackrel{~}{X}\mathrm{\#}k\overline{}_2`$ and $`X\mathrm{\#}(j+k)\overline{}_2`$ are homeomorphic, but have different invariants $`_r`$.
Proof. The simply connected $`4`$-manifolds $`\stackrel{~}{X}\mathrm{\#}k\overline{}_2`$ and $`X\mathrm{\#}(j+k)\overline{}_2`$ are non-spin and have the same invariants $`b_\pm `$; Freedmanโs classification therefore tells us that they are homeomorphic. On the other hand, because $`p_g0`$, these complex surfaces have Kodaira dimension $`0`$. Theorem 4.3 therefore tells us that
$$\frac{1}{8\pi ^2}_r(\stackrel{~}{X}\mathrm{\#}k\overline{}_2)=c_1^2(\stackrel{~}{X})+k=c_1^2(X)j+k$$
whereas
$$\frac{1}{8\pi ^2}_r(X\mathrm{\#}(j+k)\overline{}_2)=c_1^2(X)+(j+k).$$
Thus these manifolds have unequal invariants $`_r`$, as claimed.
Note that pairs $`(X,\stackrel{~}{X})`$ of the above kind are as common as garden weeds; cf. e.g. . For example, if $`X`$ is of general type, we can always find a simply connected properly elliptic surface $`\stackrel{~}{X}`$ with the same geometric genus as $`X`$.
## 5 Sectional Curvature and Volume
In the previous section, we considered a differential-topological invariant arising out of the $`L^2`$-norm of Ricci curvature. We will now turn to a related problem which arises in connection with the $`L^{\mathrm{}}`$ norm of sectional curvature.
Recall that the sectional curvature $`K(P)`$ of a Riemannian $`n`$-manifold $`(M,g)`$ is a smooth function on the Grassmann bundle of $`2`$-dimensional subspaces $`P`$ of $`TM`$. At each point $`xM`$, let us consider the bottom sectional curvature, given by
$$\underset{ยฏ}{K}(x):=\underset{PT_xM}{\mathrm{min}}K(P).$$
Then $`\underset{ยฏ}{K}:M`$ is automatically a Lipschitz continuous function, although it need not be differentiable in general. Notice that we tautologically have
$$\underset{ยฏ}{K}\frac{s}{n(n1)},$$
so that $`\underset{ยฏ}{K}`$ is negative on $`\{xM|s(x)<0\}`$.
###### Lemma 5.1
Let $`(M,g)`$ be an oriented Riemannian $`4`$-manifold, and, for $`x`$ in $`M`$, let $`w(x)0`$ denote the smallest eigenvalue of $`W_+(x):\mathrm{\Lambda }_x^+\mathrm{\Lambda }_x^+`$. Then
$$\underset{ยฏ}{K}(x)\frac{s(x)}{12}+\frac{w(x)}{2}.$$
If equality holds at $`x`$, then $`W_{}(x)=0`$; and the converse also holds if $`g`$ is Einstein.
Proof. First recall that, in terms of the decomposition
$$\mathrm{\Lambda }^2=\mathrm{\Lambda }^+\mathrm{\Lambda }^{},$$
the curvature operator of an oriented Riemannian $`4`$-manifold can be expressed in the form
=(
+W+s12BB+W-s12 ),
+W+s12BB+W-s12 {\cal R}=\left(\mbox{
\begin{tabular}[]{c|c}&\\
$W_{+}+\frac{s}{12}$&$B$\\
&\\
\cline{1-2}\cr&\\
$B^{*}$&$W_{-}+\frac{s}{12}$\\
&\\
\end{tabular}
}\right), (19)
where $`W_+`$ and $`W_{}`$ are trace-free, and where $`B:\mathrm{\Lambda }^{}\mathrm{\Lambda }^+`$ exactly corresponds to $`\stackrel{}{r}`$. On the other hand, every 2-form $`\psi `$ on $`M`$ can be uniquely written as $`\psi =\psi ^++\psi ^{}`$, where $`\psi ^\pm \mathrm{\Lambda }^\pm `$. Now a 2-form is expressible as a simple wedge product of 1-forms iff $`\psi \psi =0`$. Thus $`\psi `$ is simple and has unit length iff $`|\psi ^+|^2=|\psi ^{}|^2=\frac{1}{2}`$. The sectional curvature in the corresponding $`2`$-plane is then
$$K(P)=\frac{s}{12}+\psi ^+,W_+(\psi ^+)+\psi ^{},W_{}(\psi ^{})+2\psi ^+,B\psi ^{}.$$
(20)
Notice, however, that the last term changes sign if $`\psi ^{}`$ is replaced by $`\psi ^{}`$; geometrically, this means that the average of the sectional curvatures of an orthogonal pair of $`2`$-planes only depends on the scalar and Weyl curvatures. Thus, there is always a $`2`$-plane $`P^{}`$ for which the sectional curvature is exactly
$$K(P^{})=\frac{s}{12}+\psi ^+,W_+(\psi ^+)+\psi ^{},W_{}(\psi ^{}).$$
Hence
$`\underset{ยฏ}{K}`$ $``$ $`{\displaystyle \frac{s}{12}}+\underset{|\psi ^+|^2=\frac{1}{2}}{inf}\psi ^+,W_+(\psi ^+)+\underset{|\psi ^{}|^2=\frac{1}{2}}{inf}\psi ^{},W_{}(\psi ^{})`$ (21)
$`=`$ $`{\displaystyle \frac{s}{12}}+{\displaystyle \frac{w}{2}}+{\displaystyle \frac{\stackrel{~}{w}}{2}},`$
where $`\stackrel{~}{w}0`$ is the lowest eigenvalue of $`W_{}`$. In particular,
$$\underset{ยฏ}{K}\frac{s}{12}+\frac{w}{2},$$
as claimed.
###### Proposition 5.2
Let $`(M,g)`$ be a compact oriented Riemannian $`4`$-manifold on which there is a solution of the Seiberg-Witten equations. Let $`c_1(L)`$ be the first Chern class of the relevant spin<sup>c</sup> structure, and let $`c_1^+`$ denote its self-dual part with respect to $`g`$. Then
$$V^{1/2}\underset{xM}{\mathrm{min}}\left(\underset{ยฏ}{K}(x)+\frac{s(x)}{12}\right)\sqrt{2}\pi |c_1^+|,$$
where $`V=\text{Vol}(M,g)`$ denotes the total volume of $`(M,g)`$.
Proof. By Proposition 2.1 we know that
$$V^{1/3}\left(_M|(\frac{2}{3}s_g+2w_g)_{}|^3๐\mu \right)^{2/3}32\pi ^2(c_1^+)^2.$$
On the other hand,
$$4inf\left(\underset{ยฏ}{K}_g+\frac{s_g}{12}\right)\left(\frac{2}{3}s_g+2w_g\right)_{}$$
at all points by virtue of $`M`$ by Lemma 5.1. Hence
$$V^{1/3}\left(_M\left|4inf\left(\underset{ยฏ}{K}_g+\frac{s_g}{12}\right)\right|^3๐\mu \right)^{2/3}32\pi ^2(c_1^+)^2,$$
and so
$$\left|inf\left(\underset{ยฏ}{K}_g+\frac{s_g}{12}\right)\right|^2V2\pi ^2(c_1^+)^2.$$
Taking the square root of this inequality then yields the claim.
Let us now consider a natural family of minimal volume invariants defined with respect to only lower bounds on curvature. Let $`M`$ be any smooth compact manifold of dimension $`n`$. By just considering large constant multiples of some given metric, one obtains metrics on on $`M`$ with sectional curvature $`K1`$, at the price of making the total volume of $`M`$ enormous. This motivated Gromov to define the Gromov minimal volume
$$\text{Vol}_K(M):=inf\left\{\text{Vol}(M,g)\right|K_g1\}$$
as a way of quantifying the degree to which some negative sectional curvature might be an inevitable feature of all possible geometry on certain manifolds. However, it is also very natural to weaken the curvature hypothesis by instead only imposing lower bounds on the Ricci or scalar curvatures. In particular, one may consider what we might call the Yamabe minimal volume of $`M`$, defined by
$$\text{Vol}_s(M):=inf\left\{\text{Vol}(M,g)\right|\frac{s_g}{n(n1)}1\},$$
where our conventions have been chosen so that, by definition,
$$\text{Vol}_K(M)\text{Vol}_s(M)0.$$
One key advantage of the latter definition is that the mature theory of the Yamabe problem allows this minimal volume to be reinterpreted as
$$\text{Vol}_s(M)=\underset{g}{inf}_M\left|\frac{s_{}}{n(n1)}\right|^{n/2}๐\mu _g=\{\begin{array}{cc}0,& Y(M)0\hfill \\ \left(\frac{|Y(M)|}{n(n1)}\right)^{n/2},& Y(M)0,\hfill \end{array}$$
where $`s_{}=\mathrm{min}(s,0)`$, and where $`Y(M)`$ is the Yamabe invariant (sigma constant) of $`M`$.
It will now be convenient to consider an interpolation between these two definitions, gotten by averaging the two curvature conditions under consideration. We thus define the mixed minimal volume of a the smooth compact manifold $`M`$ to be
$$\text{Vol}_{K,s}(M)=inf\left\{\text{Vol}(M,g)\right|\frac{1}{2}\left(K_g+\frac{s_g}{n(n1)}\right)1\}.$$
This definition has been chosen in such a way that we now have the following result:
###### Theorem 5.3
Let $`M`$ be the underlying $`4`$-manifold of any complex surface of general type. Then
$$\text{Vol}_{K,s}(M)\frac{9}{4}\text{Vol}_s(M).$$
Moreover, equality holds if $`M`$ is any complex-hyperbolic manifold $`_2/\mathrm{\Gamma }`$, and in this case both minimal volumes are achieved by appropriate constant multiples of the standard metric.
Proof. Let $`X`$ be the minimal model of $`M`$. Then for every metric $`g`$ on $`M`$ there is a spin<sup>c</sup> structure with $`(c_1^+)^2c_1^2(X)`$ and such that the corresponding Seiberg-Witten invariant is non-zero for the polarization determined by $`g`$; in particular, the scalar curvature $`s`$ of $`g`$ is negative somewhere. Thus, Proposition 5.2 tells us that $`g`$ satisfies
$$V\left[\underset{xM}{\mathrm{min}}\left(\underset{ยฏ}{K}(x)+\frac{s(x)}{12}\right)\right]^22\pi ^2(c_1^+)^22\pi ^2c_1^2(X).$$
Restricting our attention to metrics for which $`\frac{1}{2}(\underset{ยฏ}{K}+\frac{s}{12})1`$, we thus get
$$\text{Vol}_{K,s}(M)\frac{2\pi ^2c_1^2(X)}{2^2}=\frac{\pi ^2}{2}c_1^2(X).$$
However , $`\text{Vol}_s(M)=\frac{2\pi ^2}{9}c_1^2(X),`$ so it follows that
$$\text{Vol}_{K,s}(M)\frac{9}{4}\text{Vol}_s(M),$$
as claimed.
If $`M`$ is a complex-hyperbolic manifold, and if $`g`$ is a multiple of the hyperbolic metric, normalized so that its sectional curvatures satisfy $`K[\frac{1}{3},\frac{4}{3}]`$, one then has $`s=8`$, and $`\frac{1}{2}(\underset{ยฏ}{K}+s)1`$, whereas $`s^2V=32\pi ^2c_1^2(M)=32(\frac{9}{2}\text{Vol}_s(M))`$. Thus $`V=\frac{9}{4}\text{Vol}_s(M)`$, and
$$\text{Vol}_{K,s}(M)=\frac{9}{4}\text{Vol}_s(M)$$
for any complex-hyperbolic $`4`$-manifold.
Because of the tautological inequality
$$\text{Vol}_K(M)\text{Vol}_{K,s}(M),$$
this immediately implies:
###### Corollary 5.4
If $`M`$ is any compact complex surface of general type, then
$$\text{Vol}_K(M)\frac{9}{4}\text{Vol}_s(M).$$
Of course, the constant appearing in this corollary should not be expected to be sharp. Indeed, inspection of the complex-hyperbolic case would instead seem to suggest the following:
###### Conjecture 5.5
If $`M`$ is any compact complex surface of general type, then
$$\text{Vol}_K(M)4\text{Vol}_s(M),$$
with equality iff $`M`$ is complex hyperbolic.
Notice that we have not even shown that only the complex hyperbolic manifolds saturate the inequality of Theorem 5.3. Nonetheless, we can at least say the following:
###### Proposition 5.6
Suppose that $`M`$ is a complex surface of general type, and suppose that the mixed minimal volume $`\text{Vol}_{K,s}(M)`$ is actually achieved by some metric $`g`$. Then either $`\text{Vol}_{K,s}(M)>\frac{9}{4}\text{Vol}_s(M)`$, or else $`(M,g)`$ is a complex-hyperbolic manifold, normalized so that $`s8`$.
Proof. Suppose that $`g`$ is a metric with $`K+\frac{s}{12}2`$ and total volume $`V=\frac{9}{4}\text{Vol}_s(M)`$. Thus, letting $`X`$ denote the minimal model of $`M`$, and choosing a spin<sup>c</sup> structure on $`M`$ as in the proof of Proposition 3.2,
$`2\pi ^2c_1^2(X)=4V`$ $``$ $`V\left[\mathrm{min}\left(\underset{ยฏ}{K}+{\displaystyle \frac{s}{12}}\right)\right]^2`$
$``$ $`{\displaystyle \frac{1}{16}}{\displaystyle (\frac{2}{3}s+2w)_{}^2๐\mu }`$
$``$ $`{\displaystyle \frac{32\pi ^2}{16}}[c_1^+]^2=2\pi ^2[c_1^+]^2`$
by virtue of (11). Since the spin<sup>c</sup> structure is chosen so that $`[c_1^+]^2c_1^2(X)`$, equality must hold, it follows that $`c_1^+=c_1(X)`$ and $`c_1^{}=0`$; in particular, $`M=X`$ is minimal. Moreover, since (11) is saturated, Proposition 2.4 tells us that $`g`$ is almost-Kรคhler, with $`s+s^{}`$ constant, $`\omega `$ an eigenvector of $`W_+`$ at each point, and $`c_1^+[\omega ]`$. Since $`c_1=c_1^+`$, the almost-complex structure satisfies $`c_1[\omega ]`$, so that our almost-Kรคhler manifold is monotonic in the terminology of .
Since equality holds in (5), we also have
$$\underset{ยฏ}{K}+\frac{s}{12}\frac{s}{6}+\frac{w}{2},$$
and this, for starters, tells us that $`W_{}0`$ by (21). Inspection of (20) then shows that $`B^{}(\omega )0`$, so that $`\omega `$ is in fact an eigenvector of the full curvature operator $``$. An almost-Kรคhler manifold with this property is called weakly $``$-Einstein. Because we also know that $`(M,g,\omega )`$ is monotonic, and that the sum $`s+s^{}`$ of the scalar and $``$-scalar curvatures is constant, a result of Drฤghici \[12, Proposition 2\] asserts that $`(M,g,J)`$ is Kรคhler-Einstein. But we also have also observed that $`W_{}0`$, so the entire curvature operator is parallel, and the universal cover of $`(M,g)`$ is therefore the symmetric space $`_2`$, equipped with the unique multiple of its standard metric for which $`s=8`$. Thus $`M`$ admits a complex-hyperbolic metric; and this metric is moreover unique, up to rescalings and diffeomorphisms, by virtue of Mostow rigidity.
We now close with some final remarks concerning minimal volumes. First of all, one might want to consider more general mixed minimal volumes, where the scalar and sectional curvatures are weighted differently from above. It may therefore be worth noting that one can also show that
$$inf\left\{\text{Vol}(M,g)\right|tK_g+(t1)\frac{s}{12}1\}(1+t)^2\text{Vol}_s(M)$$
for any complex algebraic surface $`M`$ and any constant $`t[0,\frac{1}{2}]`$; and again, equality holds for complex-hyperbolic manifolds. Indeed, it is not hard to establish inequalities such as
$$\underset{g}{inf}_M\left((1\frac{2t}{3})s+4tw\right)_{}^2๐\mu 32\pi ^2(c_1^+)^2$$
for $`t[0,\frac{1}{2}]`$, and the assertion then follows. The interesting question, though, is whether such inequalities still hold for some $`t>\frac{1}{2}`$.
It would also be interesting to determine whether
$$\text{Vol}_r(M)=inf\left\{\text{Vol}(M,g)\right|r3g\}$$
is strictly larger than $`\text{Vol}_s(M)`$ when $`M`$ is a non-minimal complex surface; our computations of $`_r(M)`$ certainly prove such an assertion for the related invariant
$$inf\left\{\text{Vol}(M,g)\right|3gr3g\}\frac{1}{36}_r(M),$$
defined with respect to two-sided curvature bounds. Because the minimizing sequences constructed in the proof of Theorem 4.3 actually have huge amounts of Ricci curvature concentrated near the blow-ups, both of these minimal volumes will presumably turn out to be much larger than $`_r/36`$ in the non-minimal case. A confirmation of this speculation would presumably also lead to even stronger non-existence theorems for Einstein metrics. And then, of course, the tautological inequality
$$\text{Vol}_K(M)\text{Vol}_r(M)$$
would most pointedly inform us that our present estimates of the Gromov minimal volume have barely begun to scratch the surface of the subject. |
warning/0003/astro-ph0003432.html | ar5iv | text | # โ
## 1 Introduction
It is agreed-upon to consider that pressure and density of a substance inside stars grow more or less monotonously approaching their centrs. It is supposed even that an ultrahigh-density substance can exist near the center of some stars. Thus, it is considered that the basic parameters of stars, such as mass, radius and temperature, are more or less arbitrary.
However, density, mass, and temperature of a star can be determined coming from the requirement of a minimum of energy if a star exists in an equilibrium state.
The effect of gravity-induced electric polarization of plasma inside a star gives a basis for such a calculation.
Under the influence of gravity, the nuclei of plasma โhangโ on their electronic clouds. Therefore, they are always located right below these clouds. As a result, electrons form a stratum at the external surface of a plasma body, and inside a star each cell of plasma obtains a positive charge $`\delta q`$. This charge is very small and equal to $`\frac{a_0}{R_{st}}e10^{19}e`$ by the order of magnitude ($`a_0`$ is Bohr radius, $`R_{st}`$ is the radius of a star, and $`e`$ is the charge of an electron). However, it is sufficient, that at each point inside a star a gravitational force is cancelled by an electric force
$$m_i\stackrel{}{g}+\delta q\stackrel{}{E}=0,$$
(1)
where $`m_i`$ is the mass of an ion, $`\stackrel{}{g}`$ is gravity acceleration, $`\stackrel{}{E}`$ is an electric field intensity created by charges $`\delta q`$
$$div\stackrel{}{E}=4\pi \delta qn,$$
(2)
where $`n`$ is the electron density.
From here it follows that the gradient of pressure inside a star is equal to zero, and the substance has a steady value of density. For a star consisting of degenerate relativistic plasma, the temperature effects can be neglected. For a star consisting of hot nonrelativistic plasma, the absence of the gradient of pressure and density of a substance requires an assumption that temperature inside a star is constant. It concerns, certainly, only the basic core of stars consisting of electrically polarized plasma. There is a thin stratum at the surface of a star, where a substance exists in an usual atomic state. In this stratum, the gradient of pressure, the gradient of density, and the gradient of temperature are present. Because of a small thickness of this stratum (about $`10^2R_{st}`$), its role in the energy balance of the whole star can be neglected.
Since the gravitational force is cancelled by the electric force in cores of stars, the gravitational energy is compensated by the electric energy inside a core as well (the gravitational energy is negative). There is an uncompensated part of energy of a gravitational field. It is the energy of gravitational field outside a star
$$E_G=\frac{GM^2}{2R},$$
(3)
where $`G`$ is the gravitational constant, $`M`$ and $`R`$ are mass and radius of a star. Because plasma is electroneutral as a whole, one proton should be related to electron of the Fermi gas of plasma. The existence of one neutron per proton is characteristic of a substance consisting of light nuclei. The quantity of neutrons increases approximately to 1.8 per proton for the heavy-nuclei substance, i.e. to one electron in plasma a mass equal to $`m_p\left(\frac{A}{Z}\right)`$ should be related. Thus
$$M=Nm_p\left(\frac{A}{Z}\right).$$
(4)
Here $`N`$ is a full number of electrons in a star, $`m_p`$ is the proton mass, $`A`$ and $`Z`$ are the mass number and charge of a nucleus, respectively. For hydrogen $`\left(\frac{A}{Z}\right)=1`$. Therefore,
$$1\left(\frac{A}{Z}\right)<2.8.$$
(5)
The field Eq.(3) tends to compress a star as a whole.
The jump in electric polarization at the surface of a star core produces additional pressure inside a star , since the jump in polarization is accompanied by the pressure jump. As this energy is considerably smaller than the gravitational one, it can be neglected for simplicity.
At high temperature the energy of the black radiation can be important in a general balance of the star energy
$$E_r=\frac{4\sigma }{c}T^4V$$
(6)
where $`\sigma =5.710^5`$ $`\frac{g}{cm^3K^4}`$ is the Stefan-Boltzmann constant, and $`V`$ is the volume of a star.
Thus, the total energy of a star is a sum of the total energy of plasma and the energy of black radiation
$$E_{total}=E_{pl}+E_r.$$
(7)
## 2 A star consisting of hot nonrelativistic electron - nuclear plasma
The energy of plasma in a star is a sum of its potential energy $`U`$ and kinetic energy $`E_{kinetic}`$. According to the virial theorem the potential energy of a system of interacting charged particles and their kinetic energy are connected
$$U=2E_{kinetic}.$$
(8)
Because the ion mass is large, the chemical potential of ions is small and their kinetic energy can be neglected. Thus, electrons will make a main contribution into the kinetic energy of plasma.
We name plasma a hot one, if the temperature inside a star is more then the degeneration temperature of the electron gas and the electric interaction between particles of plasma can be neglected
$$\left(\frac{kT}{e^2}\right)^3>>n.$$
(9)
Under this condition it is possible to consider the electron gas of plasma an ideal one, and write its equation of state as the ideal gas law
$$PV=NkT.$$
(10)
Thus, the kinetic energy of plasma in a star
$$E_{kinetic}=\frac{3}{2}NkT,$$
(11)
where $`k`$ is the Boltzmann constant.
Because gravity does not influence photons and acts on heavy ions only, the total energy of hot plasma is
$$E_{pl}=\frac{GM^2}{2R}+\frac{3}{2}NkT,$$
(12)
or, according to the virial theorem (Eq.(8)), the energy of plasma in a star
$$E_{pl}=\frac{GM^2}{4R}=\frac{3}{2}NkT$$
(13)
and the total energy of a hot star is
$$E_{total}=\frac{GM^2}{2R}+\frac{3}{2}NkT+\frac{4\sigma }{c}T^4V=\frac{GM^2}{4R}+\frac{4\sigma }{c}T^4V$$
(14)
## 3 Calculation of the corrections to the electron energy
This equation is valid for an electron gas at very high temperature, when the identity of particles is of no significance. At finite temperature the identity of particles leads to an extra stiffness of the electron gas than it follows from Eq.(10). On the other hand, electron gas in plasma will exhibit more softness under the influence of a Coulomb field of nuclei. We shall carry out the account of these corrections according to the methods described by Landau and Lifshits .
### 3.1 The correction on electron identity
The total energy of the electron gas is known
$$E_e=_0^{\mathrm{}}\epsilon ๐N_\epsilon =\frac{2^{1/2}Vm^{3/2}}{\pi ^2\mathrm{}^3}_0^{\mathrm{}}\frac{\epsilon ^{3/2}d\epsilon }{e^{(\epsilon \mu )/kT}+1},$$
(15)
where $`\epsilon =P^2/2m`$, $`P`$ and $`m`$ are momentum and mass of an electron in the nonrelativistic gas.
According to the definition of a chemical potential of electrons ,
$$\mu =kTln\left[\frac{N}{2V}\left(\frac{2\pi \mathrm{}^2}{mkT}\right)^{3/2}\right],$$
(16)
and at high temperature
$$e^{\mu /kT}<<1.$$
(17)
Therefore, in this case the integral in Eq.(15) can be expanded into the power series $`e^{\mu /kT\epsilon /kT}`$.
Keeping the first terms of series, an approximated expression of the free energy of the hot electron gas with the account of the correction on the identity of electrons is obtained
$$F=F_{ideal}+N\frac{\pi ^{3/2}e^3a_0^{3/2}n}{4(kT)^{1/2}}.$$
(18)
### 3.2 The estimation of the influence of nuclei on the hot electron gas
The nuclei reduce pressure of electron gas apart from them, therefore, this correction to pressure must be negative. The estimation of this correlative correction can be carried out by the method offered by Debye and Hรผckel .
The energy of an charged particle inside plasma is equal to $`e\phi _e`$, where $`e`$ is a charge of a particle, and $`\phi _e`$ is the electric potential created by other particles on this particle. This potential in plasma is determined by the Debye law
$$\phi =\frac{e}{r}e^{\frac{r}{r_D}}.$$
(19)
As in dense plasma nuclei form an ordered lattice , electrons play a role in screening charges only.
Thus, the Debye radius is
$$r_D=\sqrt{\frac{kT}{4\pi e^2n}}.$$
(20)
At the small value of ratio $`\frac{r}{r_D}`$, the potential can be expanded into the series
$$\phi =\frac{e}{r}\frac{e}{r_D}+\mathrm{}$$
(21)
The following terms are converted into zero at $`r=0`$. The first term of this series is the potential of the electron. The second term is the potential created by other particles on electron. Therefore, the energy of the electron gas in plasma created by the action of screened fields of other particle:
$$E_{corr}=\frac{e^2}{r_D}N=2e^3\sqrt{\frac{\pi }{kT}}\frac{N^{3/2}}{V^{1/2}}.$$
(22)
By using of the known Gibbs-Helmholtz equation
$$\frac{E}{T^2}=\frac{}{T}\frac{F}{T},$$
(23)
we obtain the correction to the free energy
$$F=F_{ideal}N\frac{2e^3}{3}\sqrt{\frac{\pi n}{kT}}.$$
(24)
### 3.3 The steady-state value of a star density
Accounting for these corrections on electron identity and influence of ions, the free energy of the hot electron gas acquires the form
$$F=F_{ideal}+N\frac{\pi ^{3/2}e^3a_0^{3/2}}{4(kT)^{1/2}}nN\frac{2e^3}{3}\frac{\pi ^{1/2}}{(kT)^{1/2}}n^{1/2}$$
(25)
From the equilibrium condition
$$\left(\frac{F}{n}\right)_{N,T}=0,$$
(26)
we obtain a steady-state value of electron density in plasma of a star
$$n_{st}=\frac{16}{9\pi ^2a_0^3}210^{23}cm^3.$$
(27)
For a star with a sufficiently high temperature, according to Eq.(9), it is energetically favourable to have this density.
### 3.4 The steady-state values of the mass, the radius and temperature of a star
According to Eq.(14), the total energy of a star
$$E_{total}\frac{3}{2}kTn_{st}V+\frac{4\sigma }{c}T^4V.$$
(28)
This function has a minimum at the temperature
$$T_{st}=\left(\frac{3ckn_{st}}{32\sigma }\right)^{1/3}=\left(\frac{ck}{6\pi ^2\sigma a_0^3}\right)^{1/3}2.510^7K.$$
(29)
According to the virial theorem Eq.(8), at an equilibrium state into the electron gas subsystem
$$\frac{GM^2}{2R}=3NkT.$$
(30)
With the account of Eq.(27) and Eq.(29), we can calculate a steady-state value of radius of hot stars
$$R_{st}=\left[\left(\frac{3}{2}\right)^2\left(\frac{\pi ck^4}{48\sigma G^3m_p^6}\right)^{1/6}\right]\frac{a_0}{\left(\frac{A}{Z}\right)}\frac{1.6}{\left(\frac{A}{Z}\right)}R_{}$$
(31)
where $`R_{}`$ is the radius of the Sun.
The steady-state value of mass of hot stars
$$M_{st}\left[\frac{27}{4}\left(\frac{c}{3\pi \sigma G^3}\right)^{1/2}\left(\frac{k}{m_p^{3/2}}\right)^2\right]\frac{m_p}{\left(\frac{A}{Z}\right)^2}\frac{1.1}{\left(\frac{A}{Z}\right)^2}M_{},$$
(32)
where $`M_{}`$ is the mass of the Sun.
## 4 The equilibrium condition of a white dwarf
When density of a substance is sufficiently high and temperature is insignificant, the electrons form a degenerate and relativistic Fermi gas. Thus, a star obtains the other equilibrium state.
It is known that the electron gas is a relativistic one, when its Fermi momentum
$$p_F=(3\pi ^2)^{1/3}n^{1/3}\mathrm{}>mc,$$
(33)
i.e., at density of particles $`n>10^{31}cm^3`$, which is characteristic for white dwarfs.
For a degeneration, electron gas temperature must be
$$T<<\frac{mc^2}{k}10^{10}K.$$
(34)
The energy of the relativistic electron gas is known
$$E=\frac{3}{4}\left(3\pi ^2\right)^{1/3}\mathrm{}cN\left(\frac{N}{V}\right)^{1/3}.$$
(35)
Therefore, the total energy of a white dwarf
$$E_{total}=\frac{GM^2}{2R}+\frac{3}{4}\left(3\pi ^2\right)^{1/3}\mathrm{}cN\left(\frac{N}{V}\right)^{1/3}.$$
(36)
As temperature is defined by the inequality Eq.(34), the temperature depending terms can be neglected. It is possible under the condition
$$\frac{4\sigma }{c}T^4V<<\frac{3}{4}\left(3\pi ^2\right)^{1/3}\mathrm{}cN\left(\frac{N}{V}\right)^{1/3}.$$
(37)
As according to the virial theorem (Eq.(8))
$$\frac{GM^2}{2R}=\frac{3}{2}\left(3\pi ^2\right)^{1/3}\mathrm{}cN\left(\frac{N}{V}\right)^{1/3},$$
(38)
the full number of electrons inside any white dwarf is fixed
$$N_{dw}=\frac{M}{\left(\frac{A}{Z}\right)m_p}=\frac{3^{5/2}\pi ^{1/2}}{2}\left(\frac{\mathrm{}c}{G\left(\frac{A}{Z}\right)^2m_p^2}\right)^{3/2}\frac{3.210^{58}}{\left(\frac{A}{Z}\right)^3}.$$
(39)
Thus, in the equilibrium state any white dwarf should have the steady-state value of mass, depending on world constants only and the ratio $`\left(\frac{A}{Z}\right)`$:
$$M_{dw}=\left[\frac{3^{5/2}\pi ^{1/2}}{2}\left(\frac{\mathrm{}c}{Gm_p^2}\right)^{3/2}\right]\frac{m_p}{\left(\frac{A}{Z}\right)^2}\frac{12.5M_{}}{\left(\frac{A}{Z}\right)^2}.$$
(40)
According to the inequality Eq.(33), the steady-state value of radius of a white dwarf
$$R_{dw}<\left(\frac{9\pi }{4}N_{dw}\right)^{1/3}\frac{\mathrm{}}{mc}310^2R_{}$$
(41)
and the density of electrons
$$n_{dw}=\frac{N_{dw}}{\frac{4\pi }{3}R_{dw}^3}>>10^{30}cm^3.$$
(42)
This correlates well with Eq.(33).
According to Eq.(37)
$$T_{dw}<<\frac{mc^{3/2}}{(4\pi )^{1/2}\sigma ^{1/4}\mathrm{}^{3/4}}310^9K.$$
(43)
Which is consistent with Eq.(34).
The steady-state value of mass of a star (40), consisting from the relativistic electron gas, does not depend on mass of an electron. Due to this fact, this equation is valid for stars, consisting of a degenerate gas of other relativistic Fermi particles, in particular, neutrons. Therefore, if we can consider a pulsar as a neutron star containing a small number of protons and electrons (with concentration more than $`10^{19}`$), the equation Eq.(40) should determine the steady-state value of mass of a pulsar.
## 5 The mass-luminosity relationship
It seems possible to assume that temperature inside a star determines temperature at its surface. If the relation between a surface temperature $`T`$ and the steady-state values of inner temperature $`T_{st}`$ is constant for all stars
$$T=constT_{st},$$
(44)
it is possible to deduce a relationship between a visible luminosity of a star and its other parameters, in particular, mass.
As the luminosity of a star
$$L=4\pi R^2\frac{\alpha }{n}T^4,$$
(45)
using the obtained above expressions and after simple transformations, we obtain
$$\frac{L}{L_{}}=\left(\frac{M}{M_{}}\right)^{10/3},$$
(46)
where $`L_{}`$ is the luminosity of the Sun.
## 6 The comparison of the calculated values with the observation data
The masses of stars can be measured with a considerable accuracy, if these stars compose a binary system. There are almost 300 double stars which masses are known with the required accuracy . Among these stars there are stars of the main sequence, dwarfs, and pulsars. Approximately one half of them are visual binaries. Their masses are measured with the a high precision. Other half consists of spectroscopic binaries and eclipsing binaries. For these stars the accuracy of mass measurement is a slightly worse. Nevertheless, we shell carry out a comparison of the calculated parameters of stars on the basis of all these measurements.
### 6.1 Masses of celestial bodies
According to these data, the distribution of masses of hot stars is described by the equality
$$<M>=(2.98\pm 0.25)M_{}.$$
(47)
It is in satisfactory agreements with the calculated steady-state value of mass of a star Eq.(32) by the order of magnitude. Graphically this distribution is shown in Fig.1. According to the binary star tables , the distribution of white dwarfs masses is described by
$$<M>=(0.96\pm 0.05)M_{}.$$
(48)
This distribution is shown in Fig. 2. In this figure, the distribution of pulsar masses is shown. It is described by
$$<M>=(1.40\pm 0.02)M_{}.$$
(49)
It is possible to conclude that with the account of the corrections on the factor $`\left(\frac{A}{Z}\right)`$, the results of calculations of the steady-state values of masses are in agreement with measured data.
### 6.2 The mass-luminosity relationship
The luminosity of stars from binary systems, depending on their masses are shown in Fig. 3. The results of more precise measurements of visual binaries are marked by squares. The data for spectroscopic and eclipsing binaries are marked by triangles and points, respectively. The line corresponds to the calculated dependence Eq.(46). It can be seen from this figure, that in the logarithmic scale the calculated dependence is in satisfactory agreement with the observed data.
## 7 The gyromagnetic ratio of stars
The effect of gravity-induced electric polarization is characteristic for all celestial bodies, consisting of a substance in a plasma state. The value of this polarization does not depend on plasma properties: no matter if it is a relativistic one or not and if it has a degeneration or not. As $`div\stackrel{}{E}=4\pi \rho `$ and $`div\stackrel{}{g}=4\pi G\gamma `$ and according to Eq.(1), the gravity-induced density of the volume electrical charge $`\rho `$ is related to the density of substance $`\gamma `$ by the relation :
$$\rho =\sqrt{G}\gamma .$$
(50)
Therefore a rotation of a celestial body about its axis must induce a magnetic field. If one assumes that the electrically polarized core of a body occupies its entire volume and if one neglects the existence of a surface stratum, where the substance is in an atomic state, that the gyromagnetic ratio (the relation of a magnetic moment of a body to its angular momentum) will obtain a steady value :
$$\vartheta =\frac{\sqrt{G}}{3c}.$$
(51)
It is possible to consider the above assumption is quite acceptable for large celestial bodies, such as stars, and less acceptable for planets, where the mantle can be rather large. However, the detailed calculation for the Earth gives the result which is in agreement with the measured value with in the accuracy of the factor of two . The values of giromagnetic ratio for all celestial bodies (for which they are known today) are shown in Fig.4.
The data for planets are taken from , the data for stars are taken from , and for pulsars - from . Therefore, for all celestial bodies - for planets and their satellites, for $`Ap`$-stars and several pulsars - the calculated value Eq.(51) with a logarithmic precision quite satisfactorily agrees to measurements, when moments itself change within the limits of more than 20 orders. |
warning/0003/hep-ph0003313.html | ar5iv | text | # Radiative Corrections to the Hadronic CrossโSection Measurement at DAฮฆNE
## 1 Introduction
The idea to use radiative events in electronโproton and electronโpositron interactions to expand the experimental possibilities for studies of different topics in high energy physics has become quite attractive in the last years. Different aspects of utilizing the radiative photons are now under intensive discussion.
Radiative events have been used already to measure the structure function $`F_2(x,Q^2)`$ at HERA . The corresponding experimental setup takes advantage of a circleโshaped photon detector (PD) in a very forward direction, as seen from the incoming electron beam. The PD measures the photon energy of all photons hitting it. The Born crossโsection for such experimental conditions has been computed in , and further theoretical study has been performed to calculate the radiative corrections (RC) to the Born crossโsection .
The possibility to undertake the $`\mathrm{{\rm Y}}`$โspectroscopy studies at the $`\mathrm{{\rm Y}}(4S)`$ energy using the emission of a hard photon from the electron or positron has been considered in . Estimates performed in this paper have demonstrated the feasibility of using the radiative photon events for the investigation of bottomonium spectroscopy at $`B`$โfactories.
Photon radiation from the initial $`e^+e^{}`$โstate in the events with missing energy has been successfully used at LEP for the measurement of the number of light neutrinos and for searches of new physics signals, see Ref. .
Recently, proposals to scan the hadronic crossโsection $`\sigma _h=\sigma (e^+e^{}hadrons)`$ at DA$`\mathrm{\Phi }`$NE energies through one such radiative process have been put forward. The strong motivation for such proposals lies in the fact that the measurement of $`\sigma _h`$, if performed below the one percent accuracy level, would allow an instructive test of the Standard Model via a precise determination of the anomalous magnetic moment $`a_\mu =(g2)_\mu /2`$ and the running electromagnetic coupling at the $`Z`$โpeak $`\alpha (M_Z^2)`$ . On-going experiments in Brookhaven will soon reduce the experimental error on $`a_\mu `$ below the precision with which the electroweak contribution to this quantity is known, and could in principle make tests of new physics. Unfortunately, this cannot yet be envisaged, since the theoretical error on $`a_\mu `$ comes mainly from the uncertainty of the hadronic vacuum polarization contribution in the energy region below and around 1 GeV, where the hadronic contribution to the photon self energy cannot be calculated unambiguously within the framework of perturbative $`QCD`$. Instead, this contribution is obtained via dispersion relations for the cross-section $`\sigma (e^+e^{}hadrons)`$ . A precise experimental determination of this quantity appears therefore the only means, at present, to reduce the theoretical error on $`a_\mu `$. There are two possible ways to measure unambiguously this cross-section in the energy region of interest, the direct scanning, as presently done at Novosibirsk , and the radiative return method. The radiative return method has a much smaller cross-section and, in order to have a statistical error in the necessary range, i.e. a fraction of a percent, it requires much more machine luminosity than the direct scanning, which is in principle the easiest to perform. Unfortunately, the reduction of the error through such direct measurement is not to be attained soon. Indeed, the precision attainable at VEPP2-M is limited by the machine luminosity, while DA$`\mathrm{\Phi }`$NE, which has a much higher design luminosity, is planning to operate at the c.m. energy $`\sqrt{s}=M_\mathrm{\Phi }`$ for the next few years. However, one can still make use of the planned DA$`\mathrm{\Phi }`$NE facility for this measurement in the near future, through the radiative return method, recently proposed as mentioned . On the theoretical side, in order to reduce the systematic errors, it is necessary to perform radiative correction to at least the percent level to the process
$$e^{}(p_1)+e^+(p_2)\gamma (k)+hadrons(q).$$
(1)
When DA$`\mathrm{\Phi }`$NE operates at the $`\mathrm{\Phi }`$โpeak, the hadronic final state is dominated by the $`\rho `$-resonance decay products and the proposals to perform the experimental scanning of the $`\pi ^+\pi ^{}\gamma `$ final state to contribute to the reduction of the error on $`a_\mu `$ has made the detailed analysis of RC to process (1) a subject of theoretical efforts.
The Born crossโsections for the radiative process of electronโpositron annihilation into a pair of charged fermions or scalar bosons were first calculated in . This topics was subsequently considered in several papers, see, for example, Refs. .
In Ref. the RC to total hadronic crossโsection of process (1) with ISR were calculated analytically for the case when the PD measures the energies of all photons emitted in the narrow cone along the direction of the electron beam. These corrections include the firstโorder contribution with the nextโtoโleading accuracy and the highโorder terms computing within the leading accuracy. At present such kind of PD is not the case for DA$`\mathrm{\Phi }`$NE. The KLOE detector allows to tag photons only outside a blind zone defined by two narrow cones along both, electron and positron beam directions. In addition, events with two hard photons tagged by the PD are rejected. Therefore, generally speaking the RC depend on โsoftโ (because of the photon energy selection in the PD) and โcollinearโ (because of the PD geometry) radiation parameters. Note that, as discussed in Section 3, the Born crossโsection depends on collinear parameters only.
An analytical calculation of the firstโorder RC to the distribution over the tagged photon energy for the KLOEโtype detector has been performed in , and an analysis of the $`\pi ^+\pi ^{}\gamma `$ final state has been carried out in using the Monte Carlo event generator for the evaluation of the RC given in .
The calculations of the RC performed in do not take into account some specific (but essential) details of the event selection in the proposed experiment with the KLOE detector . In this paper we calculate the distribution over the hadronic invariant mass in process (1) in the Born approximation and compute analytically the RC to this distribution accounting for the cuts discussed in Section 2.
## 2 Event selection in KLOE
The KLOE detector allows to measure independently the energy of the photon $`\omega `$ with the calorimeters (QCAL and EMCAL) and the invariant mass of the charged pions $`q^2`$ with the drift chamber (DC). The strategy of the experiment will be based on the measurements of the $`q^2`$ of the two pions with the DC which indirectly allows to reconstruct $`\omega `$. The much higher accuracy of the DC measurements as compared to the finite resolution of the electromagnetic calorimeter (EMCAL) is the basis for such a strategy. An attractive advantage of this approach is that, in principle, it does not require corrections of the measured distributions for the effects, caused by the experimental resolution of the calorimeter, the soโcalled deconvolution procedure, see Ref. .
Let us define the total 4โmomentum of the initial electron and positron as
$$p_1+p_2=(2E,\stackrel{}{P}_\mathrm{\Phi }),$$
where $`E`$ is the beam energy, $`\stackrel{}{P}_\mathrm{\Phi }`$ is the momentum of the $`\mathrm{\Phi }`$ and $`|\stackrel{}{P}_\mathrm{\Phi }|`$ = 12.5 MeV in the $`X`$โdirection . Note that in the laboratory frame the Lorentz boost of the $`\mathrm{\Phi }`$ is accounted for. In the interaction point the electron and positron exercise not exactly a head-on collision but there is a small beam crossing angle of order $`|\stackrel{}{P}_\mathrm{\Phi }|/2E`$ relative to the $`Z`$ axis and
$$p_1=(E,\frac{|\stackrel{}{P}_\mathrm{\Phi }|}{2},0,P_z),p_2=(E,\frac{|\stackrel{}{P}_\mathrm{\Phi }|}{2},0,P_z),P_z=E\left(1\frac{|\stackrel{}{P}_\mathrm{\Phi }|^2}{8E^2}\right).$$
(2)
Here we define the $`XZ`$ as the $`(\stackrel{}{p}_1,\stackrel{}{p}_2)`$ โ plane, and $`Z`$ as the symmetry axis of the PD.
In spite of its smallness, the quantity $`|\stackrel{}{P}_\mathrm{\Phi }|/E`$ should be taken into account in a high precision determination of the photon energy and in the calculation of the crossโsection of process (1).
In the single photon emission events the photon energy $`\omega `$ can be reconstructed directly from $`q^2=(p_1+p_2k)^2`$
$$\omega =\frac{4E^2|\stackrel{}{P}_\mathrm{\Phi }|^2q^2}{2(2E|\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi )},$$
(3)
where $`\theta (\phi )`$ is polar (azimuthal) angle of a photon in the laboratory frame. We see that because of the difference between the laboratory and the centre-of-mass frames one can reconstruct the photon energy in (1) only if the exact angular orientation of the emitted photon is known.
A systematic error could arise due to the events with the multiple photon emission. In order to reject these events as well as to decrease the background caused by the finalโstate radiation the following event selection cuts are imposed
$$\mathrm{\Omega }|\stackrel{}{K}|\eta ,\omega \omega _m,\omega _m=50MeV,\eta =10MeV,$$
(4)
where $`\mathrm{\Omega }`$ ($`\stackrel{}{K}`$) is the energy (3โmomentum) of all emitted photons, assuming that only one hard photon with the energy $`\omega `$ is tagged by the PD. Here $`\omega _m`$ denotes threshold energy for this photon. These restrictions are based on the predictions of the Monte Carlo events generator described in Ref. .
The first inequality in (4) is the reduced form of the constraint
$`M_\mathrm{\Phi }E_+E_{}|\stackrel{}{P}_\mathrm{\Phi }\stackrel{}{p}_+\stackrel{}{p}_{}|<\eta ,`$ (4a) where $`M_\mathrm{\Phi }`$ is the mass of $`\mathrm{\Phi }`$โmeson and $`E_{+,}(\stackrel{}{p}_{+,})`$ is the energy (3โmomentum) of $`\pi ^+,\pi ^{}`$ for $`\pi ^+\pi ^{}+n\gamma `$ events with $`n1.`$
Photon tagging by the QCAL calorimeter, which surrounds the blind zone and covers the angles from $`\theta _m`$ up to $`20^o`$ with respect to the electron beam direction as well as the symmetrical angles along the positron beam, can be done for the photon energy above the threshold $`\omega _{min}^{qc}=1MeV`$ (here $`\theta _m`$ is the aperture of the blind zone ). The corresponding threshold for the EMCAL calorimeter, which covers the angles from $`20^o`$ up to $`40^o`$ with respect to both the electron and the positron beam directions, is $`\omega _{min}^{ec}=5MeV)`$ . As we noted above, the events with the two hard photons inside the PD are assumed to be rejected. Therefore, when calculating the RC, one has to take into account the possibility that a soft photon with the energy $`\omega _1`$ (additional to the tagged one) hits the detector, but is not registered. Thus, we can write the following constraints on the energy $`\omega _1`$ and the radiation angle $`\theta _1`$ of an additional soft photon inside the detector
$$\omega _1<\mathrm{\Delta }_1E,if160^{}<\theta _1<\pi \theta _mand\theta _m<\theta _1<20^{},\mathrm{\Delta }_1=\frac{\omega _{min}^{qc}}{E}0.210^2;$$
(5)
$$\omega _1<\mathrm{\Delta }_2E,if40^{}>\theta _1>20^{},and160^{}>\theta _1>140^{},\mathrm{\Delta }_2=\frac{\omega _{min}^{ec}}{E}10^2,$$
where $`\theta _m`$ is for about $`10^{}`$ and $`\theta _1`$ is defined relative the Z axis.
In the following Section, we shall compute the distribution over the hadronic invariant mass in process (1) in the Born approximation.
## 3 Born approximation
To lowest order in $`\alpha `$, the differential crossโsection for process (1) with respect to the tagged hard photon has been calculated in , and here we reproduce the expression for an arbitrary hadronic final state.
The general formula for the differential crossโsection in the Born approximation can be written as
$$d\sigma ^B=\frac{2\pi ^2\alpha ^2}{S}|M|^2\frac{\alpha }{4\pi ^2}\frac{d^3k}{\omega }d\mathrm{\Gamma },$$
(6)
where $`\alpha `$ is the electromagnetic coupling , $`S=2(p_1p_2)`$ and
$$d\mathrm{\Gamma }=(2\pi )^4\delta (qq_f)\frac{d^3q_f}{2\epsilon _f(2\pi )^3}$$
is the phase space factor for the final hadrons, $`q_f`$ is the 4โmomentum of an individual hadron. The squared matrix element on the rightโhand side of Eq.(6) can be written in terms of the electronic and hadronic tensors $`L_{\mu \nu }^^\gamma `$ and $`H_{\mu \nu }`$ as
$$|M|^2=\frac{4}{q^4}L_{\mu \nu }^^\gamma H_{\mu \nu }.$$
(7)
The subscript $`\gamma `$ in the electronic current tensor indicates that here we are dealing with ISR events in process (1).
The differential cross section for radiative events can be obtained by integrating over all hadronic final states. This can be performed by using the well known relation
$$\underset{h}{}H_{\mu \nu }(q)๐\mathrm{\Gamma }=F_h(q^2)\stackrel{~}{g}_{\mu \nu },\stackrel{~}{g}_{\mu \nu }=g_{\mu \nu }\frac{q_\mu q_\nu }{q^2},$$
(8)
where the function $`F_h(q^2)`$ carries all the information about the nonโradiative hadronic crossโsection $`\sigma _h(q^2)`$. For the case of annihilation into a charged pion pair
$$F_h(q^2)=\frac{q^2|F_\pi (q^2)|^2}{24\pi }\left(1\frac{4\mu ^2}{q^2}\right)^{^{3/2}},$$
(9)
where $`\mu `$ is the pion mass and $`F_\pi (q^2)`$ is the pion electromagnetic form factor.
The leptonic tensor can be presented as
$$L_{\mu \nu }^^\gamma =\frac{(S+T_1)^2+(S+T_2)^2}{T_1T_2}\stackrel{~}{g}_{\mu \nu }+\frac{4q^2}{T_1T_2}(\stackrel{~}{p}_{1\mu }\stackrel{~}{p}_{1\nu }+\stackrel{~}{p}_{2\mu }\stackrel{~}{p}_{2\nu }),\stackrel{~}{p}=p\frac{pq}{q^2}q,$$
(10)
where we introduced the following notations
$$T_1=2p_1k=\omega \left(2E2P_z\mathrm{cos}\theta |\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi \right),$$
$$T_2=2p_2k=\omega \left(2E+2P_z\mathrm{cos}\theta |\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi \right),$$
$$S=2p_1p_2=4E^2|\stackrel{}{P}_\mathrm{\Phi }|^2,q^2=S+T_1+T_2.$$
In the expression for the leptonic tensor we have neglected terms of relative order $`m^2/|T_1|`$ and $`m^2/|T_2|`$ (here $`m`$ is the electron mass) because for the KLOE detector these cannot exceed $`m^2/(E^2\theta _m^2)10^4`$.
Taking into account that
$$L_{\mu \nu }^^\gamma \stackrel{~}{g}_{\mu \nu }=2\frac{(S+T_1)^2+(S+T_2)^2}{T_1T_2}$$
we can present the Born crossโsection as (see also Ref. )
$$d\sigma ^B=\sigma (q^2)\frac{\alpha }{2\pi ^2}\frac{(S+T_1)^2+(S+T_2)^2}{T_1T_2}\frac{d^3k}{S\omega },\sigma (q^2)=\frac{\pi \alpha ^2|F_\pi (q^2)|^2}{3q^2}\left(1\frac{4\mu ^2}{q^2}\right)^{^{\frac{3}{2}}}.$$
(11)
Let us multiply the rightโhand side of Eq.(11) by
$$dq^2\delta (4E(E\omega )|\stackrel{}{P}_\mathrm{\Phi }|^2q^2+2\omega |\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi )$$
and use the $`\delta `$โfunction to perform the integration over $`d\omega .`$ Imposing the threshold restriction (4) for the events with the emission of a single photon
$$\omega \omega _m$$
we arrive at
$$\frac{d\sigma ^^B}{dq^2}=\frac{\alpha }{2\pi ^2}\sigma (q^2)\frac{(Sq^2)d\mathrm{cos}\theta d\phi }{4S(2E|\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi )^2}\frac{(S+T_1)^2+(S+T_2)^2}{T_1T_2}$$
(12)
$$\mathrm{\Theta }(\frac{Sq^2}{2(2E|\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta \mathrm{cos}\phi )}\omega _m).$$
In principle one can perform the angular integration on the rightโhand side of Eq.(12) numerically. The analytical integration is complicated because of the $`\mathrm{\Theta }`$โfunction. This results in nontrivial limits for the angular integration. To derive them let us first examine the quantity
$$D=\frac{4E(E\omega _m)q^2|\stackrel{}{P}_\mathrm{\Phi }|^2}{2\omega _m|\stackrel{}{P}_\mathrm{\Phi }|}.$$
If $`D>1`$ then the emission of a photon is allowed in all available angular phase space. For
$$1>D>\mathrm{sin}\theta _m$$
two options appear, namely
$$2\pi >\phi >0,\mathrm{arcsin}D>\theta >\theta _m,\pi \theta _m>\theta >\pi \mathrm{arcsin}D$$
(13)
and
$$\mathrm{arccos}\frac{D}{\mathrm{sin}\theta }>\phi >0,2\pi >\phi >2\pi \mathrm{arccos}\frac{D}{\mathrm{sin}\theta },\pi \mathrm{arcsin}D>\theta >\mathrm{arcsin}D.$$
(14)
When $`\mathrm{sin}\theta _m>D>\mathrm{sin}\theta _m`$ or $`\mathrm{sin}\theta _m>D>1`$ the limits for the azimuthal integration are the same as in Eq.(14) but the polar angles are different
$$\mathrm{sin}\theta _m>D>\mathrm{sin}\theta _m,\pi \theta _m>\theta >\theta _m,$$
(15)
whereas at
$$\mathrm{sin}\theta _m>D>1,\pi \mathrm{arcsin}(D)>\theta >\mathrm{arcsin}(D).$$
(16)
Single photon emission with energy $`\omega >\omega _m`$ is not allowed if
$$D<1.$$
Using now the angular constraints given by Eqs.(13)โ(16) one can perform the angular integration in Eq.(12) analytically (at least over the azimuthal angle). The result can be presented in the following form
$$\frac{d\sigma ^^B}{dq^2}=\frac{d\sigma ^^B(D>1)}{dq^2}+\frac{d\sigma _a^^B(1>D>s_m)}{dq^2}+\frac{d\sigma _r^^B}{dq^2}.$$
(17)
The quantity $`d\sigma ^^B(D>1)/dq^2`$ corresponds to events with $`D>1`$ when all radiation angles for the tagged photon are allowed. It reads
$$\frac{d\sigma ^^B(D>1)}{dq^2}=\frac{\alpha }{2\pi }\frac{\sigma (q^2)}{2E^2}\{(\frac{S}{Sq^2}1+\frac{Sq^2}{2S})[2\mathrm{ln}\frac{1+c_m}{1c_m}+\frac{|\stackrel{}{P}_\mathrm{\Phi }|^2}{2E^2}(\mathrm{ln}\frac{1+c_m}{1c_m}+\frac{c_m}{s_m^2})]$$
(18)
$$\frac{Sq^2}{S}c_m[1+\frac{|\stackrel{}{P}_\mathrm{\Phi }|^2}{8E^2}(3c_m^2)]\},$$
where $`c_m=\mathrm{cos}\theta _m,s_m=\mathrm{sin}\theta _m.`$ The quantity $`d\sigma _a^^B(1>D>s_m)/dq^2`$ describes events with $`1>D>s_m`$ for which full coverage in the azimuthal angle is allowed. It has a structure which is very close to the rightโhand side of Eq.(18) and reads
$$\frac{d\sigma _a^^B(1>D>s_m)}{dq^2}=\frac{\alpha }{2\pi }\frac{\sigma (q^2)}{2E^2}\{(\frac{S}{Sq^2}1+\frac{Sq^2}{2S})[2\mathrm{ln}\frac{(1+c_m)(1c_d)}{(1c_m)(1+c_d)}+$$
(19)
$$\frac{|\stackrel{}{P}_\mathrm{\Phi }|^2}{2E^2}(\mathrm{ln}\frac{(1+c_m)(1c_d)}{(1c_m)(1+c_d)}+\frac{c_m}{s_m^2}\frac{c_d}{s_d^2})]\frac{Sq^2}{S}(c_mc_d)[1+\frac{|\stackrel{}{P}_\mathrm{\Phi }|^2}{8E^2}(3c_m^2c_d^2c_mc_d)]\},$$
where $`s_d=D,c_d=\sqrt{1D^2}.`$
The contribution of the remaining regions (see Eqs.(14), (15) and (16)) is described by the quantity $`d\sigma _r^^B/dq^2.`$ We integrate it only over the azimuthal angle and arrive at
$$\frac{d\sigma _r^^B}{dq^2}=\frac{\alpha }{\pi }\sigma (q^2)d\mathrm{cos}\theta [(\frac{S}{Sq^2}1+\frac{Sq^2}{2S})\frac{2}{\pi P_z\mathrm{cos}\theta }(\mathrm{\Phi }_{}\mathrm{\Phi }_+)$$
(20)
$$\frac{Sq^2}{\pi S(4E^2|\stackrel{}{P}_\mathrm{\Phi }|^2\mathrm{sin}^2\theta )}(\frac{|\stackrel{}{P}_\mathrm{\Phi }|\sqrt{\mathrm{sin}^2\theta s_d^2}}{2E|\stackrel{}{P}_\mathrm{\Phi }|s_d}+4E\mathrm{\Phi })]\mathrm{\Theta }_r,$$
where
$$\mathrm{\Phi }_\pm =\frac{1}{\sqrt{(2E\pm 2P_z\mathrm{cos}\theta )^2|\stackrel{}{P}_\mathrm{\Phi }|^2\mathrm{sin}^2\theta }}\mathrm{arctan}\sqrt{\frac{(2E\pm 2P_z\mathrm{cos}\theta +|\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta )(\mathrm{sin}\theta +s_d)}{(2E\pm 2P_z\mathrm{cos}\theta |\stackrel{}{P}_\mathrm{\Phi }|\mathrm{sin}\theta )(\mathrm{sin}\theta s_d)}}.$$
Note that on the rightโhand side of Eq.(20) $`\mathrm{sin}\theta `$ always exceeds $`s_d.`$ The quantity $`\mathrm{\Phi }`$ can be obtained from $`\mathrm{\Phi }_+`$ (or $`\mathrm{\Phi }_{}`$) by
$$\mathrm{\Phi }=\mathrm{\Phi }_\pm (P_z=0).$$
The function $`\mathrm{\Theta }_r`$ on the rightโhand side of Eq.(20) is introduced to define the upper limits of the variable $`\mathrm{cos}\theta `$ at different values of $`D.`$ It can be written explicitly as
$$\mathrm{\Theta }_r=\theta (c_d\mathrm{cos}\theta )[\theta (1D)\theta (Ds_m)+\theta (s_mD)\theta (D+1)]+$$
(21)
$$\theta (c_m\mathrm{cos}\theta )\theta (s_mD)\theta (s_m+D)$$
provided that the minimal value of $`\mathrm{cos}\theta `$ equals to zero. One can verify that in the limit $`|\stackrel{}{P}_\mathrm{\Phi }|=0`$ when only the region $`D>1`$ contributes (all emission angles are allowed) the rightโhand side of Eq.(17) coincides with the well known expression, see, for example, Refs.
$$\frac{d\sigma ^B}{dx}=\frac{\alpha }{2\pi }\sigma (q^2)2\left[\frac{1+(1x)^2}{x}\mathrm{ln}\frac{1+c_m}{1c_m}xc_m\right],$$
(22)
where in this limit $`x=\omega /E=(4E^2q^2)/(4E^2)`$ .
Note that an account for the $`|\stackrel{}{P}_\mathrm{\Phi }|/E`$ effects is mainly essential for the reconstruction of the tagged photon energy (see Eq.(3)) if one wishes to guarantee the one percent accuracy level.
## 4 Radiative corrections
The proposed high accuracy measurement of the pion contribution to the hadronic crossโsection at DA$`\mathrm{\Phi }`$NE by using radiative events in process (1), requires an adequately high precision of the theoretical predictions. These have to take into account at least the firstโorder QED radiative corrections. The firstโorder RC to $`d\sigma ^B/dq^2`$ include the real and virtual soft photon contributions in the overall phase space as well as the hard real contribution from the region where the PD does not tag a photon. Since the effect caused by the deviation of the laboratory frame from the centre-of-mass frame is small ( of relative order $`|\stackrel{}{P}_\mathrm{\Phi }|/E`$) it may be neglected in the calculation of the RC. Within this approximation we define in this section the invariants
$$s=S(|\stackrel{}{P}_\mathrm{\Phi }|=0),t_{1,2}=T_{1,2}(|\stackrel{}{P}_\mathrm{\Phi }|=0).$$
### 4.1 Virtual and soft corrections
The RC due to virtual photon emission can be computed employing the results of Ref. (see also Ref. ) where the oneโloop corrected Compton tensor with a heavy photon has been calculated for the scattering channel. In oder to obtain the corresponding results for the annihilation channel it is sufficient to make the substitutions
$$p_2p_2,us,st_2,tt_1$$
in all formulae of Ref. .
In accordance with Ref. the contribution to the differential crossโsection for process (1) due to virtual and soft photon emission can be written as
$$d\sigma ^{^{V+S}}=\frac{\alpha ^2}{8\pi ^3}\sigma (q^2)\left[\rho L_{\mu \nu }^^\gamma +T_{\mu \nu }\right]\stackrel{~}{g}_{\mu \nu }\mathrm{\Theta }\left(\frac{sq^2}{4E}\omega _m\right)\frac{d^3k}{s\omega },$$
(23)
where the quantity $`\rho `$ absorbs all infrared singularities. It can be presented as a sum of two contributions
$$\rho =\rho _V+\rho _S,$$
where $`\rho _V`$ arises due to oneโloop virtual corrections and $`\rho _S`$ โ due to soft photon contributions. For the quantity $`\rho _V`$ we can use an expression derived in
$$\rho _V=4\mathrm{ln}\frac{\lambda }{m}(L_s1)L_s^2+3L_q+\frac{4\pi ^2}{3}\frac{9}{2},L_s=\mathrm{ln}\frac{s}{m^2},L_q=\mathrm{ln}\frac{q^2}{m^2}.$$
(24)
Concerning the quantity $`\rho _S`$, the results of Ref. are not valid , since in our case the experimental requirements for the softness of an additional photon inside the PD depend on its polar angle, see Eq.(5). Note that the parameters $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$ in (5) are the physical ones and they will appear explicitly into the final expression for the RC. If an additional soft photon is outside the PD we can use an arbitrary small parameter $`\mathrm{\Delta }`$ to define its maximum energy fraction. This parameter is an auxiliary one, and it disappears in the final result for the RC due to the possibility of an additional untagged hard photon emission outside the PD (see below).
When evaluating the soft photon corrections we present the corresponding crossโsection in the factorized form
$$d\sigma ^S=d\sigma ^B\delta ,\delta =\frac{\alpha }{4\pi ^2}\frac{d^3k_1}{\omega _1}\left(\frac{p_1}{p_1k_1}\frac{p_2}{p_2k_1}\right)^2,\omega _1=\sqrt{(\stackrel{}{k}_1^2+\lambda ^2)},$$
(25)
where $`k_1`$ is the 4โmomentum of an additional soft photon. Such factorized form is valid if $`\sigma (q^2)`$ is a smooth function of $`q^2`$ , see Refs. . In the case under consideration, the width of the $`\rho `$โresonance is large enough, and approximation (25) is justified.
One can verify, whether restriction (4) $`(\mathrm{\Omega }|\stackrel{}{K}|<\eta )`$ affects this form. Obviously, Eq.(25) is valid in the case of a very soft additional radiated photon. It is, therefore, sufficient to examine its impact for the maximum allowed energy of an additional soft photon. According to Eq.(5) this maximum energy is $`\mathrm{\Delta }_2E`$. The aboveโmentioned restriction can be presented as
$$(\omega +\omega _1\eta )^2<\omega ^2+\omega _1^2+2\omega \omega _1\overline{c},$$
where $`\overline{c}`$ is the cosine of the angle between the vectors $`\stackrel{}{k}`$ and $`\stackrel{}{k}_1.`$ Since for the collinear photons $`\gamma (k)`$ and $`\gamma (k_1)`$ the constraint (4) is always fulfilled, one can check its validity for the maximal angle or for $`\overline{c}=1.`$ Setting $`\overline{c}=1`$ and $`\omega _1=\mathrm{\Delta }_2E`$ in the previous equation we obtain
$$(2\omega \eta )(2\mathrm{\Delta }_2E\eta )<0.$$
(26)
Therefore, we have
$$\eta >2\mathrm{\Delta }_2E.$$
From Eqs.(4) and (5) it follows that this restriction is satisfied. Thus, we can use representation (25) in all angular phase space for an additional soft photon.
Integration of Eq.(25) with constraints (5) and
$$\theta _1<\theta _m,\theta _1>\pi \theta _m,\omega _1<\mathrm{\Delta }E$$
leads to the following result for the RC due to soft photon emission
$$\delta =\frac{\alpha }{2\pi }\rho _S,$$
(27)
$$\rho _S=\left[4(1L_s)\mathrm{ln}\frac{\lambda }{\mathrm{\Delta }m}+L_s^2\frac{2\pi ^2}{3}+4\left(\mathrm{ln}\frac{\mathrm{\Delta }_1}{\mathrm{\Delta }}\mathrm{ln}\frac{1+c_m}{1c_m}+\mathrm{ln}\frac{\mathrm{\Delta }_2}{\mathrm{\Delta }_1}\mathrm{ln}\frac{1+c_1}{1c_1}+\mathrm{ln}\frac{\mathrm{\Delta }}{\mathrm{\Delta }_2}\mathrm{ln}\frac{1+c_2}{1c_2}\right)\right],$$
where $`c_1=\mathrm{cos}20^{}`$ and $`c_2=\mathrm{cos}40^{}`$ for the KLOE photon detector. Therefore, for the factor $`\rho `$, which is the sum of (24) and (27) we have
$$\rho =4(L_s1)\mathrm{ln}\mathrm{\Delta }+3L_q+\frac{2\pi ^2}{3}\frac{9}{2}+4\left(\mathrm{ln}\frac{\mathrm{\Delta }_1}{\mathrm{\Delta }}\mathrm{ln}\frac{1+c_m}{1c_m}+\mathrm{ln}\frac{\mathrm{\Delta }_2}{\mathrm{\Delta }_1}\mathrm{ln}\frac{1+c_1}{1c_1}+\mathrm{ln}\frac{\mathrm{\Delta }}{\mathrm{\Delta }_2}\mathrm{ln}\frac{1+c_2}{1c_2}\right).$$
(28)
Note that quantity $`\rho `$ coincides with the wellโknown expression in the limiting case $`\mathrm{\Delta }_i=\mathrm{\Delta },i=1,2`$ (see e.g. ).
Tensor $`T_{\mu \nu }`$ on the rightโhand side of Eq.(23) has the structure
$$T_{\mu \nu }=T_g\stackrel{~}{g}_{\mu \nu }+T_{11}\stackrel{~}{p}_{1\mu }\stackrel{~}{p}_{1\nu }+T_{22}\stackrel{~}{p}_{2\mu }\stackrel{~}{p}_{2\nu }T_{12}\stackrel{~}{p}_{1\mu }\stackrel{~}{p}_{2\nu }T_{21}\stackrel{~}{p}_{2\mu }\stackrel{~}{p}_{1\nu }.$$
(29)
Now after carrying out the contraction of tensors on the rightโhand side of Eq.(27) we arrive at
$$\frac{d\sigma ^{^{V+S}}}{dq^2}=\frac{\alpha }{2\pi ^2}\sigma (q^2)\frac{(sq^2)d\mathrm{cos}\theta d\phi }{4s^2}\mathrm{\Theta }(\frac{sq^2}{4E}\omega _m)$$
(30)
$$\times \frac{\alpha }{2\pi }\left[\rho \frac{(s+t_1)^2+(s+t_2)^2}{t_1t_2}+T\right],$$
$$T=\frac{3}{2}T_g\frac{1}{8q^2}\left[T_{11}(s+t_1)^2+T_{22}(s+t_2)^2+(T_{12}+T_{21})(s(s+t_1+t_2)t_1t_2)\right],s=4E^2.$$
As has been already mentioned for the KLOE detector $`|t_{1,2}|>>m^2`$, and therefore one can neglect all terms proportional to $`m^2`$ in the expressions for $`T_g`$ and $`T_{ik}.`$ Then we obtain
$$T_g=[\frac{sq^2}{t_2^2}+\frac{2s(s+t_2)+t_2^2}{t_1t_2}]G+s(\frac{q^2}{t_1t_2}\frac{2}{t_1+t_2})(L_qL_s)+\frac{s+t_1}{t_2}(\frac{3s}{s+t_2}1)(L_qL_1)+$$
(31)
$$\frac{s^2t_2^2}{2t_1t_2}+(t_1t_2),$$
$$T_{11}=\frac{2}{t_1t_2}\{q^2(1+\frac{s^2}{t_2^2})Gq^2(2+\frac{(s+t_2)^2}{t_1^2})\stackrel{~}{G}+2q^2[\frac{(s+t_2)^2}{t_1t_2}+\frac{2s}{t_1+t_2}](L_qL_s)+$$
$$\frac{4}{t_1+t_2}[s^2(s+t_2)t_1]\left[\frac{q^2}{t_1+t_2}(L_qL_s)1\right]+\frac{q^2(s+t_2)^2}{t_1(s+t_1)^2}(2s+3t_1)(L_qL_2)+$$
$$\frac{q^2}{t_2}(2st_2)(L_qL_1)4s2q^2+t_1\frac{(s+t_2)^2}{s+t_1}\},$$
(32)
$$T_{22}=T_{11}(t_1t_2,G\stackrel{~}{G}),$$
(33)
$$T_{12}+T_{21}=\frac{2}{t_1t_2}\{\frac{q^2}{t_2^2}(s+t_1)(st_2)G+\frac{q^2}{t_1^2}(sq^2t_1t_2)\stackrel{~}{G}2q^2(\frac{sq^2}{t_1t_2}+\frac{2st_2+t_1}{t_1+t_2})(L_qL_s)$$
$$\frac{4[s^2(s+t_1)t_2}{t_1+t_2}\left[\frac{q^2}{t_1+t_2}(L_qL_s)1\right]+\frac{q^2}{(s+t_1)^2}(2s+3t_1)\left(t_2\frac{q^2s}{t_1}\right)(L_qL_2)$$
$$\frac{q^2(s+t_1)}{t_2(s+t_2)}(2st_2)(L_qL_1)+8s+3t_1t_2+\frac{2st_2}{s+t_1}\}+(t_1t_2),$$
(34)
where the following notation has been introduced
$$G=(L_qL_s)(L_q+L_s2L_1)+2[f(1)+f(1\frac{q^2}{s})f(1\frac{t_1}{q^2})],\stackrel{~}{G}=G(12),$$
$$L_1=\mathrm{ln}\frac{t_1}{m^2},f(x)=\underset{0}{\overset{x}{}}\frac{dt}{t}\mathrm{ln}(1t).$$
The Bornโlike contribution (which is proportional to $`\rho `$) on the rightโhand side of Eq.(30) absorbs all infrared singularities via the quantities $`\mathrm{ln}\mathrm{\Delta },\mathrm{ln}\mathrm{\Delta }_1`$ and $`\mathrm{ln}\mathrm{\Delta }_2.`$ Concerning the collinear ones, the Bornโlike term, being integrated over the angular acceptance of the KLOE electromagnetic calorimeter, generates a contribution proportional to $`\mathrm{ln}\theta _m`$ while the remaining $`T`$-term โ to both $`\mathrm{ln}\theta _m`$ and $`(\mathrm{ln}\theta _m)^2.`$ This can be easily seen from studying the asymptotic behaviour of the term given in the second line on the rightโhand side of Eq.(30), for instance, at small values of $`|t_1|.`$ Neglecting the $`|\stackrel{}{P}_\mathrm{\Phi }|/E`$ effects we obtain in this limiting case
$$|t_1|=2\omega E(1c)E^2\theta _m^2s,|t_2|,t_2=4E^2x,q^2=4E^2(1x),c=\mathrm{cos}\theta .$$
(35)
Then we have
$$\frac{\alpha }{2\pi }\{2\rho \frac{1+(1x)^2}{x^2(1c)}+\frac{2}{x(1c)}[\frac{1+(1x)^2}{x}(\mathrm{ln}(1x)\mathrm{ln}\frac{x^2(1c)}{2(1x)}2f(x))+\frac{2x^2}{2x}]\}.$$
### 4.2 Radiation of an untagged hard photon outside the PD
When calculating RC due to the radiation of an additional invisible hard photon, we have to distinguish between the large $`(140^o>\theta _1>40^o)`$ and small $`(\theta _m>\theta _1`$ and $`\theta _1>\pi \theta _m)`$ angle radiation. For large angles we take into account only the contribution proportional to $`\mathrm{ln}\mathrm{\Delta }`$ and write it in the form
$$\frac{d\sigma ^L}{dq^2}=\frac{d\sigma ^B}{dq^2}\frac{\alpha }{2\pi }4\mathrm{ln}\frac{1}{\mathrm{\Delta }}\mathrm{ln}\frac{1+c_2}{1c_2},$$
(36)
where the Born crossโsection is defined by Eq.(12) with $`|\stackrel{}{P}_\mathrm{\Phi }|=0.`$
To simplify the calculation of the smallโangle contribution we use the quasireal electron approximation . Of course all the experimental constraints for the event selection should be taken into account.
We begin with the general expression for the cross section describing the radiation of an untagged hard photon inside the smallโangle blind zone
$$\frac{d\sigma ^^H}{dq^2}=2\frac{d\sigma _{sh}^^B}{dq^2}\frac{\alpha E^2}{4\pi ^2}\left[\frac{1+(1z)^2}{(k_2p_1)}\frac{m^2z(1z)}{(k_2p_1)^2}\right]dzd\mathrm{cos}\theta _2d\phi _2\mathrm{\Theta }_\eta ,$$
(37)
where $`z`$ is the energy fraction of an untagged hard photon $`z=\omega _2/E,`$
$$\frac{d\sigma _{sh}^^B}{dq^2}=\frac{d\sigma ^^B(p_1(1z),k,p_2)}{dq^2}=\frac{\alpha }{\pi }\sigma (q^2)\frac{s(1z)q^2}{[2E(2z(1c))]^2}$$
$$\times \frac{(1z)^2(s+t_1)^2+((1z)s+t_2)^2}{(1z)t_1t_2}\frac{d\mathrm{cos}\theta }{(1z)s}\mathrm{\Theta }[\frac{s(1z)q^2}{2E(2z(1c))}\omega _m],c=\mathrm{cos}\theta $$
(38)
is the shifted Born crossโsection (with the substitution $`p_1(1z)p_1`$) and
$$\mathrm{\Theta }_\eta =\mathrm{\Theta }(|\stackrel{}{k}+\stackrel{}{k}_2|(\omega +\omega _2\eta ))$$
is the reduced form of the restriction
$$\mathrm{\Omega }|\stackrel{}{K}|<\eta $$
for the case of one untagged hard photon with the 4โmomentum $`k_2=(\omega _2,\stackrel{}{k}_2).`$ Factor 2 on the rightโhand side of Eq.(38) appears because Eq.(37) describes collinear radiation of an additional hard photon along both electron and positron directions.
The $`\mathrm{\Theta }`$โfunction on the rightโhand side of Eq.(38) defines the maximum possible energy fraction $`z_{max}`$ of an additional hard untagged photon emitted into the smallโangle blind zone. It depends on the pion invariant mass $`q^2`$ and the polar angle $`\theta `$ of the hard photon hitting PD. Therefore it can be rewritten in the following form
$$\mathrm{\Theta }\left[\frac{s(1z)q^2}{2E(2z(1c))}\omega _m\right]=\mathrm{\Theta }(z_{max}z),z_{max}=\frac{sq^24E\omega _m}{s2E\omega _m(1c)}.$$
(39)
The $`\mathrm{\Theta }_\eta `$โfunction on the rightโhand side of Eq.(37) leads to the nontrivial correlations between the limits for variables $`z,\theta _2`$ and $`\phi _2.`$ (Here we use the coordinate frame where $`Z`$โaxis is the electron direction and define the $`(\stackrel{}{p}_1,\stackrel{}{k})`$ plane as $`XZ`$ ). These limits can be understood in following way. First one needs to analyze the quantity
$$B=\frac{z(1c_2)E\eta }{zs_2E},$$
where $`s_2=\mathrm{sin}\theta \mathrm{sin}\theta _2,c_2=\mathrm{cos}\theta \mathrm{cos}\theta _2.`$ If
$$B>1$$
then all azimuthal angles for an untagged hard photon are allowed, and we face two options for its polar angle and the energy fraction
$$I_aregion:0<\theta _2<\theta _m,\mathrm{\Delta }<z<[z_{max},\frac{\eta }{E(1c_+)}],$$
(40)
$$I_bregion:0<\theta _2<\mathrm{arccos}(1\frac{\eta }{zE})\theta ,\frac{\eta }{E(1c_+)}<z<[z_{max},\frac{\eta }{E(1c)}].$$
(41)
where $`c_+=\mathrm{cos}(\theta +\theta _m)`$ and $`[a,b]`$ is min($`a`$,$`b).`$
If the value of the quantity $`B`$ corresponds to
$$1>B>1,$$
then not all azimuthal angles for an untagged photon are allowed. In this case we obtain the following constraint
$$0<\phi _2<\mathrm{arccos}B,2\pi >\phi _2>2\pi \mathrm{arccos}B.$$
(42)
The region defined by Eq.(42) is symmetric relative to the plane $`(Z,X),`$ which contains the momentum of the photon hitting the PD . For this case there are also two possibilities for the limits for $`\theta _2`$ and $`z`$
$$II_aregion:0<\mathrm{arccos}(1\frac{\eta }{zE})\theta <\theta _2<\theta _m,\frac{\eta }{E(1c_+)}<z<[z_{max},\frac{\eta }{E(1c)}],$$
(43)
$$II_bregion=0<\theta \mathrm{arccos}(1\frac{\eta }{zE})<\theta _2<\theta _m,\frac{\eta }{E(1c)}<z<[z_{max},\frac{\eta }{E(1c_{})}],$$
(44)
where $`c_{}=\mathrm{cos}(\theta \theta _m).`$
Considering the integration limits defined by the relations (39)โ(44) one can see that there is only one region $`I_a`$ in which the untagged photon energy fraction can reach its minimal value $`\mathrm{\Delta }E.`$ Because in this region all angles for the untagged photon are allowed we can perform an angular integration on the rightโhand side of Eq.(37). The result reads
$$\frac{d\sigma ^^H(I_a)}{dq^2}=\frac{d\sigma ^^B}{dq^2}\frac{\alpha }{2\pi }\left[(4\mathrm{ln}\mathrm{\Delta }+3)L_m+4\mathrm{ln}\mathrm{\Delta }\right]+\frac{d\sigma _1^^H}{dq^2},$$
(45)
$$\frac{d\sigma _1^^H}{dq^2}=\underset{0}{\overset{[z_{max},\frac{\eta }{E(1c_+)}]}{}}2\frac{d\sigma _{sh}^^B}{dq^2}\frac{\alpha }{2\pi }[P_1(1z,L_m)2A(1z)]๐z,L_m=\mathrm{ln}\frac{E^2\theta _m^2}{m^2},$$
where
$$P_1(x,L)=lim\mathrm{\Delta }0\left[\frac{1+x^2}{1x}\theta (1x\mathrm{\Delta })+\left(\frac{3}{2}+2\mathrm{ln}\mathrm{\Delta }\right)\delta (1x)\right]L,$$
$$A(x)=lim\mathrm{\Delta }0\frac{x}{1x}\theta (1x\mathrm{\Delta })+\mathrm{ln}\mathrm{\Delta }\delta (1x).$$
In Eq.(45) we separate the dependence of the contribution caused by the untagged hard photon emission on an auxiliary parameter $`\mathrm{\Delta }.`$ One can check explicitly that this term together with (36) cancels the $`\mathrm{\Delta }`$โdependence of the the soft and virtual contribution(see Eqs.(27) and (30)).
We can also perform the analytical angular integration for the contribution of the region $`I_b`$ on the rightโhand side of Eq.(37)
$$\frac{d\sigma ^^H(I_b)}{dq^2}=\underset{\frac{\eta }{E(1c_+)}}{\overset{[z_{max},\frac{\eta }{E(1c)}]}{}}2\frac{d\sigma _{sh}^^B}{dq^2}\frac{\alpha }{2\pi }\left[\frac{1+(1z)^2}{z}\mathrm{ln}(1+\gamma )\frac{2(1z)\gamma }{z(1+\gamma )}\right]๐z,$$
(46)
$$\gamma =\frac{E^2}{m^2}\left[\mathrm{arccos}\left(1\frac{\eta }{zE}\right)\theta \right]^2.$$
Concerning the contribution of region $`II`$ in Eq.(37), we can perform the corresponding analytical integration over the azimuthal angle only. For the remaining variables $`(z,\theta _2)`$ we will show the limits of integration
$$\frac{d\sigma ^^H(II)}{dq^2}=\left\{\underset{\frac{\eta }{E(1c_+)}}{\overset{[z_{max},\frac{\eta }{E(1c)}]}{}}๐z\underset{\mathrm{arccos}\left(1\frac{\eta }{zE}\right)\theta }{\overset{\theta _m}{}}๐\theta _2+\underset{\frac{\eta }{E(1c)}}{\overset{[z_{max},\frac{\eta }{E(1c_{})}]}{}}๐z\underset{\theta \mathrm{arccos}\left(1\frac{\eta }{zE}\right)}{\overset{\theta _m}{}}๐\theta _2\right\}$$
(47)
$$\times \frac{2}{\pi }\mathrm{arccos}(B)\frac{d\sigma _{sh}^^B}{dq^2}\frac{\alpha }{2\pi }E^2\mathrm{sin}\theta _2[\frac{1+(1z)^2}{(k_2p_1)}\frac{m^2z(1z)}{(k_2p_1)^2}].$$
Thus, the contribution in the RC due to the additional untagged hard photon emission is given by the sum of Eqs. (36), (45), (46) and (47).
## 5 Full firstโorder radiative correction
The firstโorder radiative correction to the crossโsection for process (1) as measured by the KLOE detector with the above realistic experimental restrictions can be written as
$$\frac{d\sigma ^{^{RC}}}{dq^2}=\frac{\alpha }{\pi }\sigma (q^2)\frac{sq^2}{4s^2}d\mathrm{cos}\theta \mathrm{\Theta }(\frac{sq^2}{4E}\omega _m\left)\frac{\alpha }{2\pi }\right\{T+\frac{(s+t_1)^2+(s+t_2)^2}{t_1t_2}$$
(48)
$$\times [3\mathrm{ln}\frac{q^2}{s}+\frac{2\pi ^2}{3}\frac{9}{2}+(3+4\mathrm{ln}\mathrm{\Delta }_1)\mathrm{ln}\frac{4}{\theta _m^2}+4\mathrm{ln}\frac{\mathrm{\Delta }_2}{\mathrm{\Delta }_1}\mathrm{ln}\frac{1+c_1}{1c_1}\mathrm{ln}\mathrm{\Delta }_2\mathrm{ln}\frac{1+c_2}{1c_2}]\}+$$
$$\frac{d\sigma _1^^H}{dq^2}+\frac{d\sigma ^^H(I_b)}{dq^2}+\frac{d\sigma ^^H(II)}{dq^2},$$
where we used the approximation $`\mathrm{cos}\theta _m=1\theta _m^2/2.`$
The total crossโsection $`\sigma (q^2)`$ of the oneโphoton annihilation process $`e^+e^{}\pi ^+\pi ^{}`$ which should be extracted from the KLOE experiment measurements, is factorized on the right side of Eq.(48). This follows from the expression for $`d\sigma _{sh}^B/dq^2`$, see Eq.(38), which enters into each term in the third line in Eq.(48). This demonstrates an evident advantage of the approach of the Ref. as compared with the scanning of the hadron crossโsection by the tagged photon energy measurements. In the latter approach the RC caused by the additional invisible hard photon radiation include with necessity some integrals over $`\sigma (q^2)`$ . These integrals arise because in this case the tagged photon energy does not define the pion invariant mass directly.
As noted above, the auxiliary parameter $`\mathrm{\Delta }`$ disappears from the master formula (48) for the radiative correction to the crossโsection of process (1). But all physical parameters, which define the event selection (namely: the โsoftnessโ parameters $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2,`$ angular parameters $`\theta _m,\theta _1`$ and $`\theta _2`$ as well as the energy threshold $`\omega _m`$ and parameter $`\eta `$) enter this formula either explicitly or via the integration limits in the third line. The differential crossโsection over the measured $`\pi ^+\pi ^{}`$โinvariant mass $`q^2`$ is given by the sum of the Born term (17) which depends on the parameter $`|\stackrel{}{P}_\mathrm{\Phi }|`$, and (48)
$$\frac{d\sigma }{dq^2}=\frac{d\sigma ^^B}{dq^2}+\frac{d\sigma ^{^{RC}}}{dq^2}.$$
(49)
## 6 Conclusion
A crucial requirement for success of the forthcoming precision studies of the hadronic crossโsection $`\sigma (e^+e^{}hadrons)`$ at DA$`\mathrm{\Phi }`$NE through the measurements of radiative events is the matching level of reliability of the theoretical expectation. This, in turn, requires a detailed knowledge of the radiative corrections corresponding to the realistic conditions of the KLOE detector.
In this paper we derive the analytical expressions for the distribution over the invariant mass of the charged pion pair corresponding to the constraints of the proposed experiment with the KLOE detector. When DA$`\mathrm{\Phi }`$NE operates at the $`\mathrm{\Phi }`$ peak just this twoโpion final hadronic state provides the dominant contribution to the ISR events due to the radiative return to the $`\rho `$ resonance. Our approach can be quite straightforwardly extended to the description of ISR events in the general case of an arbitrary hadronic final state.
Our formulae take into account both the kinematical constraints related to the geometry of the photon detector and the event selection cuts imposed in order to reduce the FSR contamination. When deriving the Born results the Lorentz boost of the $`\mathrm{\Phi }`$ in the laboratory frame was accounted for. First of all, such an accuracy is essential for the highโprecision determination of the tagged photon energy. For the purposes of calculation of the RC with the one per cent accuracy the $`|\stackrel{}{P}_\mathrm{\Phi }|/2E`$ effects may be neglected.
A prospective advantage of the experimental strategy proposed in is the direct precise determination of the twoโpion invariant mass which, in turn allows to avoid the deconvolution procedure. One may even think about taking full advantage of the high precision of the measurements of the two charged pions with the drift chamber by making the photon tagging redundant <sup>1</sup><sup>1</sup>1We are grateful to G. Venanzoni who has attracted our attention to such an option.. An obvious attractiveness of such an inclusive strategy is that the โinvisibleโ ISR photons are then emitted dominantly in the very forward cones along the beams and the corresponding crossโsections are large (due to $`\mathrm{ln}E^2/m^2`$ enhancement). The overall event geometry becomes rather simple and the corresponding RC are governed by quasireal kinematics, see . Then the constraints imposed by the performance of KLOE calorimeters become unimportant.
The corresponding results for the Born crossโsection were presented in Ref. . The derivation of the RC requires some modifications (as compared to the results given in Ref. ) due to the contribution from additional hard photon radiation, since, in this case, the invariant mass of the pions and not the energy of the tagged photon is measured.
However, the success of such an inclusive approach requires a special care regarding different background events. Thus, a carefully chosen event selection should be introduced in order to reduce as much as possible various contaminations such as FSR events, $`\mathrm{\Phi }\pi ^+\pi ^{}\pi ^0;\pi ^+\pi ^{}\gamma `$ etc as well as doubleโphoton mechanism of $`\pi ^+\pi ^{}`$ production. Further detailed examination of the background caused by the strong decay modes and especially by the contribution of doubleโphoton $`\pi ^+\pi ^{}`$ production has to be performed. We plan to perform these studies elsewhere.
Acknowledgements
Authors thanks V.S. Fadin, G. Venanzoni and W. Kluge for fruitful discussion and critical remarks. N.P.M. thanks INFN for the hospitality. V.A.K. thanks the Leverhulme Trust for a Fellowship. This work was partly supported by the EU Framework TMR programme, contract FMRXโCT98โ0194 (DG 12โMIHT) and CTโ98โ0169.
References
1. S. Aid et al., H1 Coll., Nucl.Phys. B 470 (1996) 3.
2. M.W. Krasny, W. Plazcek, H. Spiesberger, Z.Phys. C 53 (1992) 687.
3. D. Bardin, L. Kalinovskaya, T. Riemann, Z.Phys. C 76 (1997) 487;
H. Anlauf, A.B. Arbuzov. E.A. Kuraev, N.P. Merenkov, JETP Lett. 66 (1997) 391, ibid 67 (1998) 305; JHEP 9810 (1998) 013; Phys.Rev. D 59 (1999) 014003.
H. Anlauf, A.B. Arbuzov, E.A. Kuraev. In โHamburg 1998/1999, Monte Carlo generators for HERA physicsโ, p.539, hepโph 9907248;
H. Anlauf, Eur. Phys. J. C 9 (1999) 69.
4. M. Benayoun, S.I. Eidelman, V.N. Ivanchenko, Z.K. Silagadze, hepโph/9910523.
5. For recent publication see, for example, L3 Collab., M. Acciarri, et al., Preprint CERNโEP/99โ129, Sept. 1999.
6. G. Cataldi, A. Denig, W. Kluge, G. Venanzoni, KLOE MEMO 195, August 13, 1999.
7. S. Spagnolo, โLa camera a drift di KLOE e prospettive e obiettivi di una nuova misura delle correzioni adroniche al g-2 muone a DA$`\mathrm{\Phi }`$NEโ, Doctorate thesis (in Italian), University of Lecce, 1999, unpublished;
S. Spagnolo, Eur. Phys. J. C 6 (1999) 637.
8. S. Eidelman, F. Jegerlehner, Z. Phys. C 67 (1995) 585;
F. Jegerlehner, โPrecision measurement of hadronic cross-sections at low energiesโ,
EURODA$`\mathrm{\Phi }`$NE Meeting, LNF, Frascati, Italy (1998);
F. Jegerlehner, โHadronic effects in $`(g2)_\mu `$ and $`\alpha _{QED}(M_Z)`$: Status and perspectives, Preprint DESY 99-007, hep-ph/9901386;
F. Jegerlehner, โSigma Hadronic and Precision Tests of the SMโ, LNF Spring School and EURODA$`\mathrm{\Phi }`$NE Collaboration Meeting, Frascati, Italy, (1999).
9. N. Cabibbo, R. Gatto, Phys. Rev. 124 (1961) 1577.
10. CDMโ2 Collaboration, R.R. Akhmetshin et al., Preprint Budker INPโ99โ10,
hepโex/9904027.
11. V.N. Baier, V.A. Khoze, Sov. Phys. JETP 21 (1965) 629, 1145.
12. G. Pancheri, Nuovo Cim. A 60, 321 (1969);
M. Greco, G. Pancheri and Y.N. Srivastava, Nucl. Phys. B101, 234 (1975).
13. M.J. Greutz, M.B. Einhorn, Phys. Rev. D 9 (1970) 2537;
G. Bonneau, F. Martin, Nucl. Phys. B 27 (1971) 381;
M. Greco, Riv. Nuovo Cim. 11 (1988) 1.
14. A.B. Arbuzov, E.A. Kuraev, N.P. Merenkov, L. Trentadue, JHEP 12 (1998) 009.
15. M. Konchatnij, N.P. Merenkov, JETP Lett. 69 (1999) 811.
16. S. Binner, J.H. Kลฑhn and K. Melnikov, Phys. Lett. B 459 (1999) 279.
17. M. Caffo, H. Czyz and E. Remiddi, Nuovo Cim. 110 A (1997) 515; Phys. Lett. B 327 (1994) 369.
18. E.A. Kuraev, N.P. Merenkov, V.S. Fadin, Sov. J. Nucl. Phys. 45 (1987) 486.
19. A.B. Arbuzov et al., Nucl. Phys. B 485 (1997) 457.
20. E.A. Kuraev and V.S.Fadin, Sov. J. Nucl. Phys. 41 (1985) 466.
21. V.N. Baier, V.S. Fadin, V.A. Khoze, Nucl. Phys. 65 (1973) 381. |
warning/0003/hep-th0003027.html | ar5iv | text | # The role of complex structures in ๐ค-symmetry
## 1 Introduction
In the last decade, $`W`$-algebras have provided a unifying landscape for various topics like integrable systems, conformal field theory, as well as uniformization of 2-dimensional gravity. They were originally discovered as a natural extension of the Virasoro algebra by Zamolodchikov and later implicitly by Drinfield and Sokolov . The latter obtained the classical $`W`$-algebras by equipping the coefficients of first order matrix differential operators with the second Gelโfand-Dickey Poisson bracket .
In the study of two dimensional conformal models in the so-called conformal gauge $`W`$-gravity can be defined as a generalization of the reparametrization invariance such that in the conformal gauge two copies of the corresponding $`w`$-algebra are obtained. Moreover, it has be found that the matrix differential operator has to be supplemented by another equation which is usually referred as the Beltrami equation .
Several attempts have been made in order to give a geometric picture of this very rich structure provided by $`w`$-algebras . Various aspects of this issue have been tackled independently by many authors and it is by now universally referred to as $`w`$-geometry . This geometry is naturally related to $`W`$-gravity.
On the one hand, the geometric structure turns out to be related to the uniformization of Riemann surfaces; for instance in a uniformization in higher dimensions was shown to be related to the Teichmรผller spaces constructed by Hitchin . However, the Beltrami differentials which naturally occur there are related to KdV flows but not to the complex structures underlying the $`w`$-geometry, in contrast to what it will be shown it the present paper.
On the other hand, Witten and Hull both pointed out that the use of Poisson bracket induces symplectic forms on certain manifolds and, in doing so, they proposed to study the role of symplectic diffeomorphisms in the construction of the $`w`$-algebras. Symplectic diffeomorphisms (or symplectomorphisms) form a class of diffeomorphisms on the cotangent bundle over the configuration space (phase space) which leave the canonical symplectic form invariant.
This type of invariance is very rich in the sense that it is the infinitesimal action of symplectomorphisms on a numerable set (may be even infinite) of very peculiar smooth changes of coordinates on the base Riemann surface which generates $`w`$-algebras as it has be shown quite recently . Similarly to the fact that the moduli space of Riemann surfaces plays an important role in 2d-conformal field theory coupled to gravity, it is expected that the same ought to hold for the $`w`$-symmetry .
Our treatment gives in an explicit way the infinitesimal mappings, so we can automatically give a set of Beltrami parametrizations of them which can represent this moduli space. These problems, mostly topological, have an important counterpart within Lagrangian Quantum Field Theory, involving locality: indeed the occurrence of a numerable set of โBeltrami equationsโ imposes a non local dependence of some fields on other ones, spoiling the fundamental requirements of local Quantum Field Theory. We do not realize the link between our reparametrizations and the gauge transformations of the flat $`SL(n,๐)`$ vector bundles canonically associated to the generalized projective structures, as proposed by Zucchini , and the embedding transformations proposed in References , but we hope that the absolute general statement of our transformations could shed some new light on the geometrical nature of $`w`$-symmetry.
Other examples of $`w`$ algebras, where the relevance of the complex structure occurs, were considered in many References . The former can be reconstructed by a partial breaking of the reparametrization invariance, while a set of consistency conditions controls the spoiling of the breaking of the symmetry under reparametrizations .
Moreover our approach differs slightly from these ones since it is grounded on a set of complex structures parametrized by Beltrami differentials, whose expressions give a geometrical meaning to quantities introduced in these References.
We stress that our philosophy relies on a deep connection between reparametrization invariance and $`w`$-symmetry, so if we want to investigate the physical aspects of this problem, a correct geometrical formulation will prove to be of great help.
In this context we consider a Quantum Extension of the theory, within the Lagrangian framework. However, we shall see that the locality (in the fields) assumption of a common Lagrangian Theory cannot be fulfilled; anyhow the dismission of this requirement will provide the mechanism for the compensation of the Quantum Anomalies and the improvement at the quantum level of physical quantities (Energy Momentum tensors) with definite holomorphic properties involving well defined complex structures.
Section 2 recalls some of the main results obtained in a previous work with emphasis on the notation and the basic symplectic geometry underlying our problem.
In Section 3 we study the geometrical properties of the spaces, whose mappings toward a common background space define the transformations leading to $`w`$ symmetry.
In Section 4 we find, within a B.R.S. framework, the $`w`$ algebra transformations.
In Section 5 we build a very general Lagrangian Quantum Field theory, and using the deep connections between reparametrization invariance and $`w`$ symmetry we improve its B.R.S. Quantum extension. An Appendix is devoted to the cohomological problems and and Spectral Sequences calculations.
## 2 Geometrical Approach
The starting point of a symplectic structure is the definition on a manifold of the canonical 1-form on $`T^{}(z,y)`$ $`\theta `$ defined in a local chart frame $`๐ฐ_{(z,y)}`$
$`\theta |_{๐ฐ_{(z,y)}}=\left[y_zdz+\overline{y}_{\overline{z}}d\overline{z}\right]`$ (2.1)
The $`(2,0)(0,2)`$ form $`\mathrm{\Omega }d\theta `$
$`\mathrm{\Omega }|_{๐ฐ_{(z,y)}}\left[dy_zdz+d\overline{y}_{\overline{z}}d\overline{z}\right]`$ (2.2)
is closed and
$`{\displaystyle _\mathrm{\Sigma }}\mathrm{\Omega }={\displaystyle _\mathrm{\Sigma }}\theta `$ (2.3)
Let us consider now a frame $`๐ฐ_{(Z,Y)}`$ : $`\theta `$ will take the form on $`T^{}(Z,Y)`$
$`\theta |_{๐ฐ_{(Z,Y)}}=\left[Y_ZdZ+\overline{Y}_{\overline{Z}}d\overline{Z}\right]`$ (2.4)
$`\mathrm{\Omega }`$ is globally defined and in $`๐ฐ_{(Z,Y)}`$ is defined as:
$`\mathrm{\Omega }|_{๐ฐ_{(Z,Y)}}=d\theta =\left[dY_ZdZ+d\overline{Y}_{\overline{Z}}d\overline{Z}\right]`$ (2.5)
If we require the invariance of the theory under diffeomorphisms, we have to impose that the local change of frame will generate a canonical transformation.
The change of charts will be canonical if in $`๐ฐ_{(z,y)}๐ฐ_{(Z,Y)}`$
$`\mathrm{\Omega }|_{๐ฐ_{(z,y)}}=\mathrm{\Omega }|_{๐ฐ_{(Z,Y)}}`$ (2.6)
this would imply:
$`\theta |_{๐ฐ_{(z,y)}}\theta |_{๐ฐ_{(Z,Y)}}=dF`$ (2.7)
$`F`$ is a function on $`๐ฐ_{(z,y)}๐ฐ_{(Z,Y)}`$. In the $`(z,Y)`$ plane we can define the function $`\mathrm{\Phi }(z,Y)`$ as:
$`d\mathrm{\Phi }(z,Y)d(F+(Y_ZZ+\overline{Y}_{\overline{Z}}\overline{Z})=(y_zdz+\overline{y}_{\overline{z}}d\overline{z})+(dY_ZZ+d\overline{Y}_{\overline{Z}}\overline{Z})`$ (2.8)
The function $`\mathrm{\Phi }(z,Y)`$ is the generating function for the change of charts and it is defined (up to a total differential) on the space $`๐ฐ_{(z,Y)}`$ and has non vanishing Hessian, $`{\displaystyle \frac{^2\mathrm{\Phi }}{zY}}`$.
In the region $`๐ฐ_{(z,Y)}`$ the differential operator takes the form
$`d=dz_z+d\overline{z}\overline{}_{\overline{z}}+dY_Z{\displaystyle \frac{}{Y_Z}}+d\overline{Y}_{\overline{Z}}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}(d_z+d_{Y_Z})`$ (2.9)
and $`d^2=0`$ will imply:
$`[_z,{\displaystyle \frac{}{Y_Z}}]=0[\overline{}_{\overline{z}},{\displaystyle \frac{}{Y_Z}}]=0`$ (2.10)
(and the c.c. commutators).
If we impose $`d^2\mathrm{\Phi }=0`$ we get the important properties:
$`\overline{}y_z=\overline{y}_{\overline{z}}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}Z={\displaystyle \frac{}{Y_Z}}\overline{Z}`$
$`{\displaystyle \frac{}{Y_Z}}y_z=Z{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}y_z=\overline{Z}`$ (2.11)
and their c.c., which yield that the mappings:
$`y_z(z,Y)=\mathrm{\Phi }(z,Y),Z(z,Y)={\displaystyle \frac{}{Y_Z}}\mathrm{\Phi }(z,Y)`$ (2.12)
are canonical. In particular for an arbitrary surface $`Y_Z=const`$ we can construct a local change of coordinates:
$`(z,\overline{z})(Z,\overline{Z})forZ(z,\overline{z})={\displaystyle \frac{}{Y_Z}}\mathrm{\Phi }(z,Y)|_{Y_Z=const}`$ (2.13)
In the following, for writing convenience, we shall choose this constant equal to zero.
Going on we can verify that:
$`\theta |_{๐ฐ_{(z,y)}}=d_z\mathrm{\Phi }(z,Y)`$ (2.14)
So $`\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}`$ takes the elementary form:
$`\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}`$ $`=`$ $`d\theta |_{๐ฐ_{(z,Y)}}=dY_Zdz{\displaystyle \frac{}{Y_Z}}{\displaystyle \frac{}{z}}\mathrm{\Phi }(z,Y)+d\overline{Y}_{\overline{Z}}dz{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}{\displaystyle \frac{}{z}}\mathrm{\Phi }(z,Y)`$ (2.15)
$`+dY_Zd\overline{z}{\displaystyle \frac{}{Y_Z}}{\displaystyle \frac{}{\overline{z}}}\mathrm{\Phi }(z,Y)+d\overline{Y}_{\overline{Z}}d\overline{z}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}{\displaystyle \frac{}{\overline{z}}}\mathrm{\Phi }(z,Y)`$
$`=`$ $`dY_Zd_zZ+d\overline{Y}_{\overline{Z}}d_z\overline{Z}=d_{Y_Z}y_zdz+d_{\overline{Y}_{\overline{Z}}}\overline{y}_{\overline{z}}d\overline{z}`$
$`=`$ $`d_Yd_z\mathrm{\Phi }(z,Y)`$
We shall now introduce some quantities which will be useful for our treatment.
Let us call
$`\lambda (z,Y)=_z{\displaystyle \frac{}{Y_Z}}\mathrm{\Phi }(z,Y)\lambda (z,Y)\mu (z,Y)=\overline{}_{\overline{z}}{\displaystyle \frac{}{Y_Z}}\mathrm{\Phi }(z,Y)`$
$`\overline{\lambda }(z,Y)=\overline{}_{\overline{z}}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}\mathrm{\Phi }(z,Y)\overline{\lambda }(z,Y)\overline{\mu }(z,Y)=_z{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}\mathrm{\Phi }(z,Y)`$
(2.16)
the above expression will take the form:
$`\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}`$ $`=`$ $`\left[dz\left(\lambda dY_Z+\overline{\lambda }(z,Y)\overline{\mu }(z,Y)d\overline{Y}_{\overline{Z}}\right)+d\overline{z}\left(\overline{\lambda }(z,Y)d\overline{Y}_{\overline{Z}}+\lambda (z,Y)\mu (z,Y)dY_Z\right)\right]`$ (2.17)
$`=`$ $`\left[\lambda (z,Y)\left(dz+\mu (z,Y)d\overline{z}\right)dY_Z+\overline{\lambda }(z,Y)\left(d\overline{z}+\overline{\mu }(z,Y)dz\right)d\overline{Y}_{\overline{Z}}\right]`$
due to the global definition of $`\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}`$ we can derive:
$`d_zZ(z,Y)=\lambda (z,Y)\left(dz+\mu (z,Y)d\overline{z}\right)`$ (2.18)
$`d_Yy_z(z,Y)=\lambda (z,Y)\left(dY_Z+{\displaystyle \frac{\overline{\lambda }(z,Y)\overline{\mu }(z,Y)}{\lambda (z,Y)}}d\overline{Y}_{\overline{Z}}\right)`$ (2.19)
(and their c.c.) which reveal a complex structure parametrized by an ordinary Beltrami multiplier $`\mu (z,Y)`$ in the $`(z,\overline{z})`$ plane by $`{\displaystyle \frac{\overline{\lambda }\overline{\mu }}{\lambda }}`$ in the $`(Y_Z,\overline{Y}_{\overline{Z}})`$ one.
So the $`Z(z,Y)`$ and $`y(z,Y)`$ coordinate systems can be defined in terms of a given $`\mu (z,Y)`$, by means of the Equations:
$`(\overline{}\mu (z,Y))Z(z,Y)=0`$ (2.20)
$`({\displaystyle \frac{}{Y_Z}}{\displaystyle \frac{\lambda (z,Y)\overline{\mu }(z,Y)}{\lambda (z,Y)}}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}})y_z(z,Y)=0`$ (2.21)
From the previous equations the Liouville theorem will follow:
$`det|{\displaystyle \frac{Z(z,Y)}{z}}|=\lambda (z,Y)\overline{\lambda }(z,Y)(1\mu (z,Y)\overline{\mu }(z,Y))=det|{\displaystyle \frac{y_z(z,Y)}{Y}}|`$ (2.22)
Using the previous parametrization, as is well known, we can write the derivative operators $`_Z,\frac{}{Y_Z}`$ as:
$`_Z={\displaystyle \frac{_z\mu (z,Y)\overline{}_{\overline{z}}}{\lambda (z,Y)(1\mu (z,Y)\overline{\mu }(z,Y))}}`$ (2.23)
$`{\displaystyle \frac{}{y_z(z,Y)}}={\displaystyle \frac{๐^z\mu (z,Y)\overline{๐}^{\overline{z}}}{1\mu (z,Y)\overline{\mu }(z,Y)}}`$ (2.24)
(and their c.c) where we have introduced:
$`๐^z(z,Y)={\displaystyle \frac{1}{\lambda (z,Y)}}{\displaystyle \frac{}{Y_Z}}`$ (2.25)
Finally, if we work in the $`๐ฐ_{(z,Y)}`$ space, taking $`z`$ and $`Y_Z`$ as passive coordinates, the condition $`d\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}=0`$ will give:
$`d\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}`$ $`=`$ $`\overline{}\lambda (z,Y)d\overline{z}dzdY_Z+(\lambda (z,Y)\mu (z,Y))dzd\overline{z}dY_Z`$ (2.26)
$`+\overline{\lambda }(z,Y)dzd\overline{z}d\overline{Y}_{\overline{Z}}+\overline{}(\overline{\lambda }(z,Y)\overline{\mu }(z,Y))d\overline{z}dzd\overline{Y}_{\overline{Z}}`$
$`+d\overline{Y}_{\overline{Z}}{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}\left(\lambda (z,Y)\left(dz+\mu (z,Y)d\overline{z}\right)\right)dY_Z`$
$`+dY_Z{\displaystyle \frac{}{Y_Z}}\left(\overline{\lambda }(z,Y)\left(d\overline{z}+\overline{\mu }(z,Y)dz\right)\right)d\overline{Y}_{\overline{Z}}=0`$
which gives rise to the Beltrami identities:
$`\overline{}\lambda (z,Y)=(\mu (z,Y)\lambda (z,Y)),{\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}\lambda (z,Y)={\displaystyle \frac{}{Y_Z}}\left(\overline{\lambda }(z,Y)\overline{\mu }(z,Y)\right)`$ (2.27)
and their c.c.
It is important to remark that from Eq (2.27) one has,
$`{\displaystyle \frac{๐^z\overline{\lambda }(z,Y)}{\overline{\lambda }(z,Y)}}={\displaystyle \frac{\overline{๐}^{\overline{z}}\mu (z,Y)+\mu (z,Y)๐^z\overline{\mu }(z,Y)}{1\mu (z,Y)\overline{\mu }(z,Y)}}`$ (2.28)
and its c.c. From Eq (2.10) we also get:
$`[_z,๐^z]`$ $`=`$ $`_z\mathrm{log}\lambda (z,Y)๐^z`$
$`[\overline{}_{\overline{z}},๐^z]`$ $`=`$ $`\overline{}_{\overline{z}}\mathrm{log}\lambda (z,Y)๐^z=\left(\mu (z,Y)๐^z\mu (z,Y)[_z,๐^z]\right)`$ (2.29)
(and their c.c.), and the commutator,
$`[๐^z,\overline{๐}^{\overline{z}}]=(๐^z\overline{\mu }(z,Y)){\displaystyle \frac{}{y_z(z,Y)}}(\overline{๐}^{\overline{z}}\mu (z,Y)){\displaystyle \frac{}{\overline{y}_{\overline{z}}(z,Y)}}`$ (2.30)
from which it follows:
$`[{\displaystyle \frac{}{y_z(z,Y)}},{\displaystyle \frac{}{\overline{y}_{\overline{z}}(z,Y)}}]=0.`$ (2.31)
At this stage, one remark is in order. In Eq (2.15) the terms $`d_{Y_Z}ZdY_Z+d_{Y_Z}\overline{Z}d\overline{Y}_{\overline{Z}}`$ and $`d_zy_zdz+d_z\overline{y}_{\overline{z}}d\overline{z}`$ will identically vanish in $`\mathrm{\Omega }`$. Accordingly, we can state the important Theorem :
###### Theorem 2.1
On the smooth trivial bundle $`\mathrm{\Sigma }\times \text{R}^2`$, the vertical holomorphic change of local coordinates,
$`Z((z,\overline{z}),(Y_Z,\overline{Y}_{\overline{Z}}))Z((z,\overline{z}),(Y_Z),\overline{Y}_{\overline{Z}}),`$ (2.32)
where $``$ is a holomorphic function in $`Y_Z`$, while the horizontal holomorphic change of local coordinates,
$`y_z(z,\overline{z},(Y_Z,\overline{Y}_{\overline{Z}}))y_z(f(z),\overline{z},(Y_Z,\overline{Y}_{\overline{Z}})),`$ (2.33)
where $`f`$ is a holomorphic function in $`z`$, are both canonical transformations.
In the first case an infinitesimal variation of $`Z(z,Y)`$ in $`Y_Z`$ does not modify, for fixed $`(z,\overline{z})`$ the $`\mathrm{\Omega }`$ form.
So the diffeomorphisms $`(z)Z((z,\overline{z}),Y_Z);zZ((z,\overline{z}),Y_Z+dY_Z)`$, will be related to the same two form $`\mathrm{\Omega }`$.
If we make the expansion around, say,$`Y_Z=0,\overline{Y}_{\overline{Z}}=0`$ the generating function $`\mathrm{\Phi }`$ will be written as the series :
$`\mathrm{\Phi }(z,Y)`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{n_{max}}{}}}{\displaystyle \frac{1}{n!}}\left[Y_Z^n\left({\displaystyle \frac{}{Y_Z}}\mathrm{\Phi }_1((z,\overline{z}),Y_Z)\right)^n|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}\right]`$ (2.34)
$`+{\displaystyle \underset{n=1}{\overset{n_{max}}{}}}{\displaystyle \frac{1}{n!}}\left[\overline{Y}_{\overline{Z}}^n\left({\displaystyle \frac{}{\overline{Y}_{\overline{Z}}}}\overline{\mathrm{\Phi }_1}((z,\overline{z}),\overline{Y}_{\overline{Z}})\right)^n|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}\right]`$
$``$ $`{\displaystyle \underset{n=1}{\overset{n_{max}}{}}}\left[Y_Z^nZ^{(n)}(z,\overline{z})\right]+{\displaystyle \underset{n}{}}\left[\overline{Y}_{\overline{Z}}^n\overline{Z}^{(n)}(z,\overline{z})\right]`$
where:
$`Z^{(n)}(z,\overline{z})\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\mathrm{\Phi }_1(z,Y)\right]|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}`$ (2.35)
And we can reconstruct $`Z(z,Y)`$ as:
$`Z(z,Y)={\displaystyle \underset{n}{}}nZ^{(n)}(z,\overline{z})Y_{Z}^{}{}_{}{}^{n1}`$ (2.36)
We shall see in the following that the family of reparametrizations:
$`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})`$ (2.37)
will be the origin of the $`w`$ algebras symmetry transformations. Obviously the choice of the $`(Y_Z,\overline{Y}_{\overline{Z}})`$ origin as starting point does not alter the treatment
The symplectic form will then be written:
$`\mathrm{\Omega }|_{๐ฐ_{(z,Y)}}=d_Yd_z\mathrm{\Phi }(z,Y)={\displaystyle \underset{n}{}}\left[d_YY_Z^nd_zZ^{(n)}(z,\overline{z})+d_Y\overline{Y}_{\overline{Z}}^nd_z\overline{Z}^{(n)}(z,\overline{z})\right]`$ (2.38)
Note that the conjugate momenta to the complex coordinates $`Z^{(n)},n1`$ are related to the n-th power of the conjugate momenta of the coordinate $`Z^{(1)}(z,\overline{z})`$.
## 3 The geometry of the $`Z^{(n)}`$-spaces
In the previous chapter we introduced the mappings:
$`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})n=1\mathrm{}n`$ (3.1)
foreseeing that they will be fundamental for our purposes.
Before investigating their role in the construction of $`w`$-algebras, we shall derive down below the most important properties of the $`(Z^{(n)},\overline{Z}^{(n)})`$-spaces.
### 3.1 Generalities on $`(Z^{(n)},\overline{Z}^{(n)})`$-spaces
The $`Z^{(n)}(z,\overline{z})`$ coordinates are defined as:
$`Z^{(n)}(z,\overline{z})=[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n(\mathrm{\Phi }(z,Y))]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$ (3.5)
$`=`$ $`{\displaystyle \underset{j=1,\mathrm{},n}{}}j!{\displaystyle \underset{i=1,\mathrm{},m_j}{}}\left[{\displaystyle \frac{(Z^{(p_i)}(z,\overline{z}))^{a_i}}{a_i!}}\right]|_{\left\{\begin{array}{c}_ia_i=j\\ _ia_ip_i=n\\ p_1>p_2>,\mathrm{},>p_{m_j}\end{array}\right\}}^{(j)}(z,\overline{z})`$
where the functions $`^j(z,\overline{z})`$ are given by:
$`^{(j)}(z,\overline{z}){\displaystyle \frac{1}{j!}}\left(๐\right)^j\mathrm{\Phi }(z,Y)|_{Y_Z,\overline{Y_Z}=0}`$ (3.6)
Note that the $`Z^{(n)}(z,\overline{z})`$ coordinate is no more independent from $`Z^{(r)}`$, for $`r=1,\mathrm{},n1`$; by the way it obeys differential consistency conditions induced by the $`^{(j)}(z,\overline{z})`$ functions: we shall sketch the above system for further convenience:
$`Z(z,\overline{z})`$ $`=`$ $`Z^{(1)}(z,\overline{z})`$
$`Z^{(2)}(z,\overline{z})`$ $`=`$ $`Z^{(2)}(z,\overline{z})^{(1)}(z,\overline{z})+(Z)^2^{(2)}(z,\overline{z})`$
$`Z^{(3)}(z,\overline{z})`$ $`=`$ $`Z^{(3)}(z,\overline{z})^{(1)}(z,\overline{z})+2Z(z,\overline{z})Z^{(2)}(z,\overline{z})^{(2)}(z,\overline{z})+Z(z,\overline{z})^3^{(3)}(z,\overline{z})`$
$`\mathrm{}`$
$`Z^{(N)}(z,\overline{z})`$ $`=`$ $`Z^{(N)}(z,\overline{z})^{(1)}(z,\overline{z})+\mathrm{}\mathrm{}\mathrm{}+\left(Z(z,\overline{z})\right)^N^{(N)}(z,\overline{z})`$ (3.7)
where the first equation can be solved by
$`\mathrm{ln}Z(z,\overline{z})={\displaystyle _{\stackrel{~}{z}}^z}๐z^{}{\displaystyle \frac{1}{^1(z^{},\overline{z})}}+\mathrm{ln}Z(\stackrel{~}{z},\overline{z})`$ (3.8)
where $`\mathrm{ln}Z(\stackrel{~}{z},\overline{z})`$ takes into account the boundary conditions. Thus, one has
$`Z(z,\overline{z})=Z(\stackrel{~}{z},\overline{z})\mathrm{exp}{\displaystyle _{\stackrel{~}{z}}^z}๐z^{}{\displaystyle \frac{1}{^{(1)}(z^{},\overline{z})}}`$ (3.9)
which plugged into the second equation gives $`Z^{(2)}`$ as an integral relation between $`Z`$, $`^{(1)}(z,\overline{z})`$ and $`^{(2)}(z,\overline{z})`$;
$`Z^{(2)}(z,\overline{z})=Z(z,\overline{z})\left[{\displaystyle ^z}๐z\mathrm{"}{\displaystyle \frac{^{(2)}(z\mathrm{"},\overline{z})Z(z\mathrm{"},\overline{z})}{(^1(z\mathrm{"},\overline{z}))^3}}\right]`$ (3.10)
In full generality, one gets,
$`Z^{(N)}(z,\overline{z})=Z(z,\overline{z})\left[{\displaystyle ^z}๐z\mathrm{"}{\displaystyle \frac{1}{Z(z\mathrm{"},\overline{z})}}^N(Z^{(j)}(^k(z\mathrm{"},\overline{z})_{kj})^i(z\mathrm{"},\overline{z}))\right]`$ (3.11)
where $`^{(i)}(z,\overline{z})i=1,2,\mathrm{},n`$ are fixed and $`Z^{(j)}(z\mathrm{"},\overline{z})j<n`$ have been calculated at the preceding orders. So we can state the following
###### Theorem 3.1
The set of functions $`^{(i)}(z,\overline{z})`$, $`i=1,\mathrm{},n`$ completely identify the set of coordinates $`Z^{(j)}(z,\overline{z})`$, $`j=1,\mathrm{},n`$
Now we want to solve another problem: given a local change of coordinates $`(z,\overline{z})(๐ต(z,\overline{z}),\overline{๐ต}(z,\overline{z}))`$ is it possible to consider it as an element of arbitrary n-th order of a $`w`$ hierarchy $`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})`$, and to find a construction of the underlying $`(Z^{(i)},\overline{Z}^{(i)})i=1,\mathrm{},n1`$ in order to get a $`w`$ description of this local change?
The answer is positive, since using a standard construction of the $`(Z^{(i)},\overline{Z}^{(i)}),i=1,\mathrm{},n1`$ spaces (which do not interfere with $`(Z^{(n)},\overline{Z}^{(n)})`$), we can use it in the last equation of (3.7) and get in an algebraic way the suitable solution $`^N(z,\overline{z})`$. So we can state the Theorem:
###### Theorem 3.2
For an arbitrary local change: $`(z,\overline{z})(๐ต(z,\overline{z}),\overline{๐ต}(z,\overline{z}))`$ it is possible to generate a space hierarchy $`(z,\overline{z})(Z^{(r)},\overline{Z}^{(r)}),r=1,\mathrm{},n`$.
Finally we explore the $`(Z^{(n)},\overline{Z}^{(n)})`$ spaces with respect to the $`(z,\overline{z})`$ background. As in , we shall introduce:
$`d_zZ^{(n)}(z,\overline{z})`$ $`=`$ $`\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\mathrm{\Phi }(z,Y)dz+{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\overline{}\mathrm{\Phi }(z,Y)d\overline{z}\right]|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}`$ (3.12)
$``$ $`\lambda _z^{Z^{(n)}}(z,\overline{z})[dz+\mu _{\overline{z}}^z(n,(z,\overline{z}))d\overline{z}]`$
where:
$`\lambda _z^{Z^{(n)}}(z,\overline{z})\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\mathrm{\Phi }(z,Y)\right]|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}Z^{(n)}(z,\overline{z})`$
$`\lambda _z^{Z^{(n)}}(z,\overline{z})\mu _{\overline{z}}^z(n,(z,\overline{z}))\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\overline{}\mathrm{\Phi }(z,Y)\right]|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}\overline{}Z^{(n)}(z,\overline{z}).`$ (3.13)
So we shall define, for all $`n`$:
$`{\displaystyle \frac{}{Z^{(n)}}}={\displaystyle \frac{\overline{\mu }(n,(z,\overline{z}))\overline{}}{(1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z})))}}`$ (3.14)
In particular we get from Eq (3.5):
$`{\displaystyle \frac{}{\overline{Z}^{(n)}}}\left({\displaystyle \underset{j=1}{\overset{n}{}}}j!{\displaystyle \underset{i=1}{\overset{m_j}{}}}\left[{\displaystyle \frac{(\lambda _z^{Z^{(p_i)}}(z,\overline{z})))^{a_i}}{a_i!}}\right]|_{\left\{\begin{array}{c}_ia_i=j\\ _ia_ip_i=n\\ p_1>p_2>,\mathrm{},>p_{m_j}\end{array}\right\}}^{(j)}(z,\overline{z})\right)=0.`$ (3.18)
This identity which now appears trivial, will acquire a particular meaning in the following.
It is so evident that the quantity $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ will label the complex structure of the space $`Z^{(n)}`$ in the $`(z,\overline{z})`$ background and increasing the order $`n`$ this complex structure will explore the complex structure of all the $`(z,Y)`$ space.
More precisely we get:
$`\lambda (z,Y)\mu (z,Y)`$ $`=`$ $`{\displaystyle \underset{n}{}}n\lambda _z^{Z^{(n)}}(z,\overline{z})\mu (n,(z,\overline{z}))Y_{Z}^{}{}_{}{}^{n1}`$
$`\lambda (z,Y)`$ $`=`$ $`{\displaystyle \underset{n}{}}n\lambda _z^{Z^{(n)}}(z,\overline{z})Y_{Z}^{}{}_{}{}^{n1}`$ (3.19)
The symplectic form will reduce to:
$`\mathrm{\Omega }|_{๐ฐ(z,Y)}={\displaystyle \underset{n=1\mathrm{}n_{max}}{}}\left[\lambda _z^{Z^{(n)}}(z,\overline{z})dY_{Z}^{}{}_{}{}^{n}+\overline{\lambda }_{\overline{z}}^{Z^{(n)}}\overline{\mu }(n,(z,\overline{z}))d\overline{Y}_{\overline{Z}}^{}{}_{}{}^{n}\right]dz`$
$`+{\displaystyle \underset{n=1\mathrm{}n_{max}}{}}\left[\overline{\lambda }_{\overline{z}}^{Z^{(n)}}(z,\overline{z})d\overline{Y}_{\overline{Z}}^{}{}_{}{}^{n}+\lambda _z^{Z^{(n)}}(z,\overline{z})\mu (n,(z,\overline{z}))dY_{Z}^{}{}_{}{}^{n}\right]d\overline{z}`$ (3.20)
From the very definition, the Beltrami parameter will take the general form:
$`\mu _{\overline{z}}^z(n,(z,\overline{z})){\displaystyle \frac{\overline{}Z^{(n)}(z,\overline{z})}{Z^{(n)}(z,\overline{z})}}={\displaystyle \underset{j=1}{\overset{n}{}}}j!{\displaystyle \underset{i=1}{\overset{m_j}{}}}\left[{\displaystyle \frac{(\lambda _z^{Z^{(p_i)}}(z,\overline{z}))^{a_i}}{a_i!\lambda _z^{Z^{(n)}}(z,\overline{z})}}\right]|_{\left\{\begin{array}{c}_ia_i=j\\ _ia_ip_i=n\\ p_1>p_2>,\mathrm{},>p_{m_j}\end{array}\right\}}\mu _{\overline{z}}^{(j)}(z,\overline{z}).`$ (3.24)
where we have introduced:
$`\mu _{\overline{z}}^{(n)}(z,\overline{z})=\left[{\displaystyle \frac{1}{(n)!}}\left(๐^z\right)^n\overline{}\mathrm{\Phi }(z,Y)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$ (3.25)
We remark that, due to Eq.(2.19) the presence of $`\lambda `$โs in the $`Y_Z`$ derivative compromises the locality requirements but the parameters in Eq (3.25) introduce a suitable parametrization for a local Lagrangian Quantum Field Theory use. Furthermore these ones have to be considered as the least common factors for all the Beltrami factors of the spaces $`Z^{(r)}(z,\overline{z})r=1,\mathrm{},n`$
Note that the Beltrami multiplier $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ will depend on the $`\lambda ^{}s`$ parameters of the spaces parametrized by the $`Z^{(i)}`$ coordinates with $`in`$.
Under a change of background coordinates the Beltrami multiplier transforms as:
$`\mu _{\overline{z}}^z(n,(z,\overline{z}))=\mu _{\overline{z}^{}}^z^{}(n,(z^{},\overline{z^{}}))(^{}z)(\overline{}\overline{z}^{})`$ (3.26)
so we can derive:
$`\mu _{\overline{z}}^{(n)}(z,\overline{z})=\mu _{\overline{z}^{}}^{(n^{})}(z^{},\overline{z^{}})(^{}z)^n(\overline{}\overline{z}^{})`$ (3.27)
giving a well-defined geometrical status to $`\mu _{\overline{z}}^n`$ as a $`(n,1)`$-conformal field.
A Beltrami identity is immediately recovered for each (and any) $`n`$ as a consequence of $`d_z^2=0`$,
$`\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\overline{}\mathrm{\Phi }(z,Y)\right]|_{Y_Z=0,\overline{Y}_{\overline{Z}}=0}=\overline{}\lambda _z^{Z^{(n)}}(z,\overline{z})=(\lambda _z^{Z^{(n)}}(z,\overline{z})\mu _{\overline{z}}^z(n,(z,\overline{z})))`$ (3.28)
It is not only obvious that $`\lambda _z^{Z^{(n)}}`$ is a non local object on $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$, but, due to Eq.(3.25) the parameters $`\lambda _z^{Z^{(n)}}`$ is not local on the parameters $`\mu _{\overline{z}}^{(i)}(z,\overline{z})`$ with $`in`$. Furthermore from Eq (3.24) we can realize that the Beltrami multiplier $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ (see Eqs(3.28) (3.25) ) is sensible to the complex structures of the inner subspaces ; so we can state:
###### Statement 3.1
a) The $`\lambda _z^{Z^{(n)}}`$ (for fixed n) are non local functions of the parameters $`\mu _{\overline{z}}^{(r)}`$, and will contain all of them with order $`rn`$.
b)The complex structure of the $`(Z^{(n)},\overline{Z}^{(n)})`$ spaces can be described by parameters $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ which extend to these spaces the Beltrami multipliers. For a given $`n`$, $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ will depend, through $`\lambda ^{Z^{(r)}},r<n`$, in a non local way on the $`\mu _{\overline{z}}^{(j)}`$โs with $`jn`$.
c) As a consequence of the previous statements and of Eq(3.12) the coordinate $`Z^{(n)}(z,\overline{z})`$ is a non local function of $`\mu _{\overline{z}}^{(j)}(z,\overline{z})`$ with $`jn`$
These are important geometrical statements of the work and are the basis for the physical discussion of the problem. So the mappings
$`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})n`$ (3.29)
are non holomorphic and the non holomorphicity depends on $`n`$.
The only possibility to get local Beltrami multipliers $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ is to have $`\mu _{\overline{z}}^{(r)}(z,\overline{z})=0r>1`$ . In this case the complex structure of all the $`(Z^{(n)},\overline{Z}^{(n)})`$ space will coincide with the $`(Z^{(1)},\overline{Z}^{(1)})`$ one. So necessarily $`r2`$:
$`\left[๐^z\right]^r\overline{}\mathrm{\Phi }(z,Y)|_{Y_Z,\overline{Y}_{\overline{Z}}=0}=\left[{\displaystyle \frac{1}{\lambda (z,Y)}}{\displaystyle \frac{}{Y_Z}}\right]^r{\displaystyle \underset{n}{}}\left[Y_Z^n\overline{}Z^{(n)}(z,\overline{z})\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}=0,`$ (3.30)
It is easy to see that the previous equation leads to:
$`{\displaystyle \frac{\overline{}Z^{(r)}(z,\overline{z})}{Z^{(r)}(z,\overline{z})}}={\displaystyle \frac{\overline{}Z(z,\overline{z})}{Z(z,\overline{z})}}\mu (z,\overline{z}),r2`$ (3.31)
which, expressed in terms of the $`(Z,\overline{Z})`$ background, gives:
$`(\overline{}\mu (z,\overline{z}))Z^{(r)}=0,r2,`$ (3.32)
showing that $`Z^{(r)}`$ is an holomorphic function of $`Z`$.
This result is straightforwardly generalized. Imposing $`rl+1`$:
$`\left[๐^z\right]^r\overline{}\mathrm{\Phi }(z,Y)|_{Y_Z,\overline{Y}_{\overline{Z}}=0}=\left[{\displaystyle \frac{1}{\lambda (z,Y)}}{\displaystyle \frac{}{Y_Z}}\right]^r{\displaystyle \underset{n}{}}\left[Y_Z^n\overline{}Z^{(n)}(z,\overline{z})\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}=0,`$ (3.33)
one recovers
$`{\displaystyle \frac{\overline{}Z^{(r)}(z,\overline{z})}{Z^{(r)}(z,\overline{z})}}={\displaystyle \frac{\overline{}Z^{(l)}(z,\overline{z})}{Z^{(l)}(z,\overline{z})}}\mu (l,(z,\overline{z})),rl+1`$ (3.34)
which leads to:
$`\overline{}_{Z^{(l)}}Z^{(r)}=0,rl+1.`$ (3.35)
Thus $`Z^{(r)}`$ is holomorphic in $`Z^{(l)}`$. Obviously the inverse is always true ; so we can state:
###### Theorem 3.3
The set of conditions: $`\mu _{\overline{z}}^{(r)}=0,r=l+1,\mathrm{},n`$ will imply that $`Z^{(n)}`$ is an holomorphic function of $`Z^{(l)}`$ and vice-versa.
### 3.2 Complex structures in $`(Z^{(r)},\overline{Z}^{(r)})`$ backgrounds
It is already interesting to analyze the complex structure of $`(Z^{(n)},\overline{Z}^{(n)})`$-space with respect to different backgrounds; indeed setting
$`dZ^{(n)}=\mathrm{\Lambda }_{Z^{(r)}}^{Z^{(n)}}(Z^{(r)},\overline{Z}^{(r)})\left[dZ^{(r)}+\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))d\overline{Z}^{(r)}\right]`$ (3.36)
where:
$`\mathrm{\Lambda }_{Z^{(r)}}^{Z^{(n)}}(Z^{(r)},\overline{Z}^{(r)}){\displaystyle \frac{Z^{(n)}(Z^{(r)},\overline{Z}^{(r)})}{Z^{(r)}}}`$
$`\mathrm{\Lambda }_{Z^{(r)}}^{Z^{(n)}}(Z^{(r)},\overline{Z}^{(r)})\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}){\displaystyle \frac{\overline{Z}^{(n)}(Z^{(r)},\overline{Z}^{(r)})}{Z^{(r)}}}`$ (3.37)
so that the quantity $`\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))`$ is the Beltrami multiplier of the coordinates $`Z^{(n)}`$ in the $`(Z^{(r)},\overline{Z}^{(r)})`$ background $`n`$. So we can relate these quantities to the corresponding objects relatively to the $`(z,\overline{z})`$ background, $`n`$ and $`r`$:
$`\mathrm{\Lambda }_{Z^{(r)}}^{Z^{(n)}}(Z^{(r)},\overline{Z}^{(r)})|_{(z,\overline{z})}={\displaystyle \frac{\lambda _z^{Z^{(r)}}(z,\overline{z})\left(1\mu (r,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))\right)}{\lambda _z^{Z^{(n)}}(z,\overline{z})\left(1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))\right)}}\overline{\mu }(n,(z,\overline{z})))`$ (3.38)
and:
$`\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))|_{(z,\overline{z})}={\displaystyle \frac{\lambda ^{Z^{(r)}}(z,\overline{z})\left(\mu (r,(z,\overline{z}))\mu (n,(z,\overline{z}))\right)}{\overline{\lambda }^{\overline{Z}^{(r)}}(z,\overline{z})\left(1\mu (r,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))\right)}}`$ (3.39)
also we derive:
$`[\mathrm{\Lambda }_{Z^{(r)}}^{Z^{(n)}}(Z^{(r)},\overline{Z}^{(r)})\overline{\mathrm{\Lambda }}_{(\overline{Z}^r)}^{(\overline{Z}^n)}(Z^{(r)},\overline{Z}^{(r)})(1\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))\overline{\mathrm{\Xi }}_{Z^{(r)}}^{\overline{Z}^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))]|_{(z,\overline{z})}`$
$`={\displaystyle \frac{\lambda _z^{Z^{(n)}}(z,\overline{z})\overline{\lambda }_{\overline{z}}^{Z^{(n)}}(z,\overline{z})(1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z})))}{\lambda ^{Z^{(r)}}(z,\overline{z})\overline{\lambda }^{\overline{Z}^{(r)}}(z,\overline{z})(1\mu (r,(z,\overline{z}))\overline{\mu }(r,(z,\overline{z})))}}`$ (3.40)
## 4 $`w`$ B.R.S algebras
The previous construction introduces to a BRS derivation of $`w`$-symmetry as shown in . Our aim is to construct a BRS differential which considers, in an infinitesimal approach all the mappings $`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})`$, for all $`n`$, on the same footing.
Consider first the infinitesimal variations $`\mathrm{\Lambda }(z,Y)`$ of the generating function $`\mathrm{\Phi }(z,Y)`$ under the diffeomorphism action of the cotangent bundle. Then by taking the expansion (2.34) one can proceed as follows . Let $`๐ฎ`$ be the BRS diffeomorphism operator acting on the $`(z,\overline{z})`$ basis and defined for each $`n`$ by
$`๐ฎZ^{(n)}(z,\overline{z})\mathrm{{\rm Y}}^{(n)}(z,\overline{z})\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\mathrm{\Lambda }(z,Y)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$ (4.1)
consequently the $`๐ฎ`$ nilpotency will give:
$`๐ฎ\mathrm{{\rm Y}}^{(n)}(z,\overline{z})=0`$ (4.2)
we shall decompose:
$`\mathrm{{\rm Y}}^{(n)}(z,\overline{z})\left[\lambda _z^{Z^{(n)}}(z,\overline{z})๐ฆ_n^z(z,\overline{z})\right]`$ (4.3)
and we get from the definition of $`\lambda _z^{Z^{(n)}}(z,\overline{z})`$ :
$`๐ฎ\lambda _z^{Z^{(n)}}(z,\overline{z})`$ $`=`$ $`\left[{\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^n\mathrm{\Lambda }(z,Y)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$ (4.4)
$``$ $`\left[\left({\displaystyle \frac{1}{n!}}\left({\displaystyle \frac{}{Y_Z}}\right)^{n1}(\lambda (z,Y)๐(z,Y))\right)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$
$`=`$ $`\left(\lambda _z^{Z^{(n)}}(z,\overline{z})๐ฆ_n^z(z,\overline{z})\right)`$
and consequently:
$`๐ฎ๐ฆ_n^z(z,\overline{z})=๐ฆ_n^z(z,\overline{z})๐ฆ_n^z(z,\overline{z})`$ (4.5)
Expanding the calculations yields:
$`๐ฆ_n^z(z,\overline{z})={\displaystyle \underset{j=1}{\overset{n}{}}}j!{\displaystyle \underset{i=1}{}}m_j\left[{\displaystyle \frac{(\lambda _z^{Z^{(p_i)}}(z,\overline{z}))^{a_i}}{a_i!\lambda _z^{Z^{(n)}}}}\right]|_{\left\{\begin{array}{c}_ia_i=j\\ _ia_ip_i=n\\ p_1>p_2>\mathrm{}>p_{m_j}\end{array}\right\}}๐^{(j)}(z,\overline{z})`$ (4.9)
where we have introduced:
$`๐^{(n)}(z,\overline{z})=\left[{\displaystyle \frac{1}{n!}}\left(๐^z\right)^n\mathrm{\Lambda }(z,Y)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0},n=1,2\mathrm{}`$ (4.10)
which, for all $`n`$, provide, in Eq (4.9), a geometric expansion with the same non local coefficients as in Eq (3.24).
It is quite easy, from the very definition, to derive that these ghosts transform as:
$`๐ฎ๐^{(n)}={\displaystyle \underset{r=1}{\overset{n}{}}}\left(r๐^{(r)}_z๐^{(nr+1)}\right)`$ (4.11)
revealing the holomorphic $`w`$ character of these ghosts. We remark that the B.R.S. variations of $`๐^{(n)}(z,\overline{z})`$ depends on the fields $`๐^{(r)}(z,\overline{z})`$ with $`rn`$. The upper limit of the indices of these ghosts coincides with the the upper index of the expansion Eq (2.34),and will characterize this symmetry: we do not fix it, and our conclusions hold their validity for any value (finite or infinite) of this index.
The connection between Eqs (4.9) and (3.24) spreads new light on the connection between diffeomorphisms and $`w`$ algebras by putting into the game the complex structures.
Now the coefficients of these expansions are essentially geometrical factors.
On the other hand the quantities $`\mu _{\overline{Z}}^{(n)}(z,\overline{z})`$ in Eq (3.25) will have the B.R.S variations:
$`๐ฎ\mu _{\overline{z}}^{(n)}(z,\overline{z})=\overline{}C^{(n)}(z,\overline{z}){\displaystyle \underset{r=1}{\overset{n}{}}}[r\mu _{\overline{z}}^{(r)}(z,\overline{z})C^{(nr+1)}(z,\overline{z})`$
$`rC^{(r)}(z,\overline{z})\mu _{\overline{z}}^{(nr+1)}(z,\overline{z})]`$ (4.12)
and:
$`๐ฎ\left[\lambda _z^{Z^{(n)}}(z,\overline{z})\mu _{\overline{z}}^z(n,(z,\overline{z}))\right]=\overline{}\mathrm{{\rm Y}}^{(p)}`$ (4.13)
so that:
$`๐ฎ\mu _{\overline{z}}^z(n,(z,\overline{z}))=๐ฆ_n^z(z,\overline{z})\mu _{\overline{z}}^z(n,(z,\overline{z}))\mu _{\overline{z}}^z(n,(z,\overline{z}))๐ฆ_n^z(z,\overline{z})+\overline{}๐ฆ_n^z(z,\overline{z})`$ (4.14)
So for each $`n`$ a diffeomorphic structure with a ghost $`๐ฆ_n^z`$ (non local in the complex structure parameter ) can be put into evidence
From Eqs (3.24) it is easy to find the form of these variations in terms of $`w`$ components. We shall introduce:
$`\kappa _n^z(z,\overline{z}){\displaystyle \frac{๐ฆ_n^z(z,\overline{z})\mu (n,(z,\overline{z}))\overline{๐ฆ_n}^{\overline{z}}(z,\overline{z})}{1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))}}`$ (4.15)
for which:
$`\mathrm{{\rm Y}}^{(n)}(Z^{(n)},\overline{Z}^{(n)})_{Z^{(n)}}+\overline{\mathrm{{\rm Y}}}^{(\overline{n})}(Z^{(n)},\overline{Z}^{(n)})_{\overline{Z}^{(n)}}=\kappa _n^z(z,\overline{z})+\overline{\kappa _n^z}(z,\overline{z})\overline{}`$ (4.16)
so:
$`๐ฆ_n^z(z,\overline{z})=\kappa _n(z,\overline{z})+\mu (n,(z,\overline{z}))\overline{\kappa }_n^{\overline{z}}(z,\overline{z})`$ (4.17)
In this base we have:
$`๐ฎ\kappa _n^z(z,\overline{z})=\kappa _n^z(z,\overline{z})\kappa _n^z(z,\overline{z})+\overline{\kappa _n}^{\overline{z}}(z,\overline{z})\overline{}\kappa _n^z(z,\overline{z})`$ (4.18)
$`๐ฎ\mu (n,(z,\overline{z}))`$ $`=`$ $`(\kappa _n^z+\overline{\kappa _n}^{\overline{z}}\overline{})\mu (n,(z,\overline{z}))\mu (n,(z,\overline{z}))(\kappa _n^z+\mu (n,(z,\overline{z}))\overline{\kappa }^{\overline{z}})`$ (4.19)
$`+\overline{}\kappa _n^z+\mu (n,(z,\overline{z}))\overline{}\overline{\kappa _n}^{\overline{z}}`$
$`๐ฎ\lambda _z^{Z^{(n)}}(z,\overline{z})=(\kappa _n^z+\overline{\kappa _n}^{\overline{z}}\overline{})\lambda _z^{Z^{(n)}}(z,\overline{z})+\lambda _z^{Z^{(n)}}(z,\overline{z})(\kappa _n^z+\mu (n,(z,\overline{z}))\overline{\kappa }^{\overline{z}})`$ (4.20)
Note that the condition:
$`๐ฎ\mathrm{{\rm Y}}^{(n)}(z,\overline{z})=0`$ (4.21)
is verified only if the Beltrami condition Eq(3.28) holds.
In this approach we can also derive:
$`๐^{(j)}(z,\overline{z})={\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s>0\end{array}}{\overset{j}{}}}\left[r!s!\left(\mathrm{\Pi }_i{\displaystyle \frac{\left(\mu _{\overline{z}}^{(l_i)}(z,\overline{z})\right)^{k_i}}{k_i!}}\right)|_{\left\{\begin{array}{c}_ik_i=s\\ r+_il_ik_i=j\end{array}\right\}}\right]c^{(r,s)}(z,\overline{z})`$ (4.26)
where we have introduced, in the spirit of Eq (2.24) the new ghosts:
$`c^{(p,q)}(z,\overline{z})=\left[{\displaystyle \frac{1}{p!}}{\displaystyle \frac{1}{q!}}\left({\displaystyle \frac{}{y_z(z,Y)}}\right)^p\left({\displaystyle \frac{}{\overline{y}_{\overline{z}}(z,Y)}}\right)^q\mathrm{\Lambda }(z,Y)\right]|_{Y_Z,\overline{Y}_{\overline{Z}}=0}`$
(4.27)
of conformal weight $`(p,q)`$ and which transform as:
$`๐ฎc^{(p,q)}(z,\overline{z})={\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s1\end{array}}{\overset{r=p,s=q}{}}}(rc^{(r,s)}(z,\overline{z})_zc^{(pr+1,qs)}(z,\overline{z})`$ (4.30)
$`+sc^{(r,s)}(z,\overline{z})\overline{}_{\overline{z}}c^{(pr,qs+1)}(z,\overline{z}))`$ (4.31)
We also remark that the variation of $`c^{(p,q)}`$ contains the ghost fields $`c^{(r,s)}`$ with lower degrees, $`rp;sq`$.
Combining together Eqs.(4.10) and (4.26) we shall write:
$`๐ฆ_n^z(z,\overline{z})={\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s1\end{array}}{\overset{n}{}}}\eta _{(r,s)}^z(n,(z,\overline{z}))c^{(r,s)}(z,\overline{z})`$ (4.34)
where
$`\eta _{(r,s)}^z(n,(z,\overline{z}))={\displaystyle \underset{j=\text{max}(r,s)}{\overset{n}{}}}j!{\displaystyle \underset{i=1}{\overset{m_j}{}}}\left[{\displaystyle \frac{(\lambda _z^{Z^{(p_i)}}(z,\overline{z}))^{a_i}}{a_i!\lambda _z^{Z^{(n)}}}}\right]`$ (4.35)
$`\times `$ $`\left[r!s!\left(\mathrm{\Pi }_i{\displaystyle \frac{\left(\mu _{\overline{z}}^{(l_i)}(z,\overline{z})\right)^{k_i}}{k_i!}}\right)\right]|_{\left\{\begin{array}{c}_ik_i=s\\ r+_il_ik_i=j\\ _ia_i=j\\ _ia_ip_i=n\\ p_1>p_2>\mathrm{}>p_{m_j}\end{array}\right\}}`$ (4.41)
in particular:
$`\eta _{(1,0)}^z(n,(z,\overline{z}))`$ $`=`$ $`1`$
$`\eta _{(0,1)}^z(n,(z,\overline{z}))`$ $`=`$ $`\mu _{\overline{z}}^z(n,(z,\overline{z}))`$ (4.42)
The same can be written in terms of the $`\kappa _n(z,\overline{z})`$ ghosts, getting:
$`\kappa _n^z(z,\overline{z})=c^{(1,0)}(z,\overline{z})`$ (4.45)
$`+{\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s2\end{array}}{\overset{n}{}}}{\displaystyle \frac{\left(\eta _{(r,s)}^z(n,(z,\overline{z}))\mu (n,(z,\overline{z}))\overline{\eta }_{(r,s)}^{\overline{z}}(z,\overline{z})\right)}{1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))}}c^{(r,s)}(z,\overline{z}).`$
### 4.1 The introduction of matter field sectors in $`w_n`$-algebras
The introduction of matter in $`w`$ invariant models, and in particular $`w`$ gravity, have been treated in the literature in different scenarios according to the different attempts to reach $`w`$ algebras, e.g .
Our point of view, which relates in a geometrical fashion $`w`$ algebras to two dimensional conformal field theory, heavily supports the methods which introduce matter in conformal models. A proper $`(\alpha ,\beta )`$-differential has to be invariant under holomorphic changes of charts, thus induces a local rescaling by the $`\lambda _z^{Z^{(n)}}`$โs,
$`\varphi _{(\alpha ,\beta )}(Z^{(n)},\overline{Z}^{(n)})(dZ^{(n)})^\alpha (d\overline{Z}^{(n)})^\beta `$ $`=`$ $`(\lambda _z^{Z^{(n)}}(z,\overline{z}))^\alpha (\overline{\lambda }_{\overline{z}}^{\overline{Z}^{(n)}}(z,\overline{z}))^\beta \varphi _{(\alpha ,\beta )}(Z^{(n)},\overline{Z}^{(n)})`$
$`\times (dz+\mu (n,(z,\overline{z}))d\overline{z})^\alpha (d\overline{z}+\overline{\mu }(n,(z,\overline{z}))dz)^\beta `$
$`\phi _{(\alpha ,\beta )}(z,\overline{z})(dz+\mu (n,(z,\overline{z}))d\overline{z})^\alpha (d\overline{z}+\overline{\mu }(n,(z,\overline{z}))dz)^\beta `$
The field will be said scalar if $`(\alpha ,\beta )=(0,0)`$, namely,
$`\varphi (Z^{(n)},\overline{Z}^{(n)})\phi (z,\overline{z}),n`$ (4.47)
and will transform with only under point displacement
$`๐ฎ\varphi (Z^{(n)},\overline{Z}^{(n)})=\left(\mathrm{{\rm Y}}^{(n)}_{Z^{(n)}}+\mathrm{{\rm Y}}^{\overline{Z}^{(n)}}\overline{}_{\overline{Z}^{(n)}}\right)\varphi (Z^{(n)},\overline{Z}^{(n)})`$ (4.48)
Going now to background $`(z,\overline{z})`$ we have
$`๐ฎ\phi (z,\overline{z})=(\kappa _n^z(z,\overline{z})+\overline{\kappa _n}^{\overline{z}}(z,\overline{z})\overline{})\phi (z,\overline{z})`$ (4.49)
Now each $`(Z^{(n)},\overline{Z}^{(n)})`$ space has a different complex structure, so using the background representation each field living in this space carries into its transformation the imprinting of this space.
$`๐ฎ\varphi _{(\alpha ,\beta )}(Z^{(n)},\overline{Z}^{(n)})`$ $`=`$ $`\left(\mathrm{{\rm Y}}^{(n)}_{Z^{(n)}}+\mathrm{{\rm Y}}^{\overline{Z}^{(n)}}_{\overline{Z}^{(n)}}\right)\varphi _{(\alpha ,\beta )}(Z^{(n)},\overline{Z}^{(n)})`$ (4.50)
$`=`$ $`(\kappa _n^z+\overline{\kappa _n}^{\overline{z}}\overline{})\phi _{(\alpha ,\beta )}(z,\overline{z})`$
The same can be done with respect to the background system of coordinates
$`๐ฎ\phi _{(\alpha ,\beta )}(z,\overline{z})`$ $`=`$ $`(\kappa _n^z+\overline{\kappa }_n^{\overline{z}}\overline{})\phi _{(\alpha ,\beta )}(z,\overline{z})`$
$`+\alpha (\kappa _n^z+\mu (n,(z,\overline{z}))\overline{\kappa }_n^{\overline{z}})\phi _{(\alpha ,\beta )}(z,\overline{z})`$
$`+\beta (\overline{}\overline{\kappa _n}^{\overline{z}}+\overline{\mu }(n,(z,\overline{z}))\overline{}\kappa _n^z)\phi _{(\alpha ,\beta )}(z^\alpha \overline{z}^\beta )(z,\overline{z})`$
In conclusion the above ghost decompositions Eqs(4.9) (4.34) clarify our strategy towards a treatment of $`w`$ algebras in a Lagrangian Quantum Field Theory framework ; since from the former it is quite straightforward to derive in combination with the canonical construction of the diffeomorphism B.R.S. operator, the one induced by the $`w`$ ordinary algebras. This will be very useful in the next Section.
The diffeomorphism variations of the โmatter fieldsโ $`\varphi _{(\alpha ,\beta )}(Z^{(n)},\overline{Z}^{(n)})`$, fix, from our point of view , their $`w`$ transformations, since it will be provided by the decomposition of ghosts $`\kappa _n^z(z,\overline{z})`$ in terms of the true $`c^{(r,s)}(z,\overline{z})`$ symplectomorphism ghost fields.
In particular, for the scalar field, the B.R.S. variation Eq(4.1) is rewritten as:
$`๐ฎ\phi (z,\overline{z})={\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s1\end{array}}{\overset{n}{}}}c^{(r,s)}(z,\overline{z})\left(\tau _{n,(r,s)}^z+\overline{\tau }_{n,(r,s)}^{\overline{z}}\overline{}\right)\phi (z,\overline{z})`$ (4.54)
We remark that this description is totally different from the the various approach to $`w`$ matter found in the literature e.g. . Moreover, according to this viewpoint, it gives completely trivial the problem.
## 5 Lagrangian Field Theory building
The approach we have given before to $`w`$ algebras, and in particular the relevance of complex structures in their construction, suggests to investigate the role played by these symmetries in a Lagrangian Field Theory.
In the previous Sections we have emphasized the linking points between $`w`$ algebras and two dimensional conformal transformations, so our discussion starts with an example of conformal invariant models.
As it is well known, two dimensional conformal symmetry means reparametrization invariance: in our $`w`$ scheme we have to improve the symmetry of a wide class of reparametrization mappings, so a lot of care is required in order to respect the Lagrangian Field Theory prescriptions.
We shall deal with a common conformal model built on a two dimensional space manifold $`๐ต(z,\overline{z}),\overline{๐ต}(z,\overline{z})`$.
In order to construct a properly defined local Lagrangian theory we have to take care of:
1) a well definition of the Action integral,
2) the locality on the constituent fields,
3) the symmetry constraints.
We consider the scalar field case:
$`\mathrm{\Gamma }_{scalar}={\displaystyle ๐๐ต_๐ต\varphi (๐ต,\overline{๐ต})}d\overline{๐ต}\overline{}_{\overline{๐ต}}\varphi (๐ต,\overline{๐ต})`$ (5.1)
$`{\displaystyle ๐Z^{(n)}_{Z^{(n)}}\varphi (Z^{(n)},\overline{Z}^{(n)})}d\overline{Z}^{(n)}\overline{}_{\overline{Z}^{(n)}}\varphi (Z^{(n)},\overline{Z}^{(n)})`$
So we shall start from a model which is invariant under a reparametrization $`(z,\overline{z})(๐ต(z,\overline{z}),\overline{๐ต}(z,\overline{z}))`$, which is well defined but has quantum anomalies.
Our strategy for the realization of a $`w`$ symmetry in this model will be to consider the $`๐ต(z,\overline{z}),\overline{๐ต}(z,\overline{z})`$ space as an $`n`$-th element of a $`w`$ space hierarchy as in Eq(3.7).
A positive answer for our purposes comes from Theorem (3.2) but more care has to be exercised.
For this reason the model is to be โwell definedโ with respect all the possible backgrounds ; indeed the Lagrangian in Eq (5.1) written in terms of the $`(z,\overline{z})`$ background takes the form:
$`\mathrm{\Gamma }_{scalar}`$ $``$ $`{\displaystyle ๐z}d\overline{z}_{z,\overline{z}}(z,\overline{z})`$ (5.2)
$`=`$ $`{\displaystyle ๐z}d\overline{z}{\displaystyle \frac{\left[\overline{\mu }(n,(z,\overline{z}))\overline{}\right]\varphi (z,\overline{z})\left[\overline{}\mu (n,(z,\overline{z}))\right]\varphi (z,\overline{z})}{\left(1\mu (n,(z,\overline{z}))\overline{\mu }(n,(z,\overline{z}))\right)}}`$
Moreover the model is well defined in each $`(Z^{(r)},\overline{Z}^{(r)})`$ frame ($`r`$) since:
$`\mathrm{\Gamma }_{scalar}`$ $`=`$ $`{\displaystyle ๐Z^{(n)}_{Z^{(n)}}\varphi (Z^{(n)},\overline{Z}^{(n)})}d\overline{Z}^{(n)}\overline{}_{\overline{Z}^{(n)}}\varphi (Z^{(n)},\overline{Z}^{(n)})`$
$`=`$ $`{\displaystyle ๐Z^{(r)}}d\overline{Z}^{(r)}\left(1\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))\overline{\mathrm{\Xi }}_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))\right)^1`$
$`\times \left[_{Z^{(r)}}\overline{\mathrm{\Xi }}_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))\overline{}_{\overline{Z}^{(r)}}\right]\varphi (z,\overline{z})\left[\overline{}_{\overline{Z}^{(r)}}\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))_{Z^{(r)}}\right]\varphi (z,\overline{z})`$
This means that in this framework we can assume as symmetry transformations the changes of charts:
$`(z,\overline{z})(Z^{(r)}(z,\overline{z}),\overline{Z}^{(r)}(z,\overline{z}))r=1\mathrm{}nn`$ (5.4)
just defined in Eq (3.7); and the dynamics of the particle, which is free and scalar in the space $`(Z^{(n)},\overline{Z}^{(n)})`$, if described by means of the background of the underlying complex spaces $`(Z^{(r)}(z,\overline{z}),\overline{Z}^{(r)}(z,\overline{z})r=1\mathrm{}n1`$ need the parametrization of the Beltrami multiplier $`\mu _{\overline{z}}^z(n(z,\overline{z}))`$ just found in the Eq (3.24).
So at the light of previous arguments and of Theorem (3.2)
###### Statement 5.1
A two dimensional conformal model admits, at the classical limit, a $`w`$-symmetry of arbitrary order.
Anyhow the quantum extension requires some care.
Indeed the $`\lambda `$โs, are non local functions of the $`\mu _{\overline{z}}^{z^s}(z,\overline{z})`$,so in a local Lagrangian Quantum Field theory approach, they are not primitive, but, they are essential for the geometrical meaning of our $`w`$ construction.
We have just seen in the preceding Lagrangian construction that, if we want to maintain the โwell definitionโ of the Lagrangian with respect all the $`(Z^{(r)},\overline{Z}^{(r)})`$ frames, they are contained in the Beltrami $`\mathrm{\Xi }_{\overline{Z}^{(r)}}^{Z^{(r)}}(n,(Z^{(r)},\overline{Z}^{(r)}))`$ due to Eqs (3.39) (3.24).
If we want to analyze how the underlying complex structure contributes to the dynamics the price to pay is to put into the game all the $`\lambda _z^{Z^{(r)}}`$ fields, $`r=1,\mathrm{},n`$ induced by the decomposition of the $`(Z^{(n)},\overline{Z}^{(n)})`$-spaces Eq.(3.5). These fields (even if local in the $`(z,\overline{z})`$ background) are non local in the $`\mu _{\overline{z}}^{(r)}(z,\overline{z})r=1\mathrm{}n`$ fields due to the Beltrami equations (3.28) in each $`(Z^{(r)},\overline{Z}^{(r)})`$, $`r=1,\mathrm{},n`$ sectors. So, according to this point of view, the model becomes intrinsically non local in the fields (unless $`n=1`$).
We have now to choose the set of fields which exhausts the dynamical configuration space : so our coordinates will be the fields: $`\phi `$, $`c^{(r,s)}`$, $`\mu _{\overline{z}}^{(r)}`$ ,$`\lambda _z^{Z^r}`$, $`r=1,\mathrm{},n`$ and their derivatives. This means that all the Classical B.R.S. diffeomorphism transformations of these fields have to be written in terms of the $`c^{(p,q)}`$ ghosts using the expansion of the various ghosts $`๐^{(j)}`$, $`๐ฆ_n^z`$ and $`\kappa _n^z`$ as written previously.
So we define as โnaive โ BRS functional operator the following $`\delta _c`$:
$`\delta _c={\displaystyle \underset{n=1}{\overset{n_{max}}{}}}{\displaystyle }dzd\overline{z}[(\lambda _z^{Z^{(n)}}(z,\overline{z})(\kappa _n^z+\mu (n,(z,\overline{z}))\overline{\kappa _n}^{\overline{z}}(z,\overline{z})){\displaystyle \frac{\delta }{\delta Z^{(n)}(z,\overline{z})}}`$
$`+{\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s1\end{array}}{\overset{n}{}}}\left(rc^{(r,s)}(z,\overline{z})_zc^{(pr+1,qs)}(z,\overline{z})+sc^{(r,s)}(z,\overline{z})\overline{}_{\overline{z}}c^{(pr,qs+1)}(z,\overline{z})\right){\displaystyle \frac{\delta }{\delta c^{(p,q)}(z,\overline{z})}}`$ (5.7)
$`+\left(\overline{}C^{(n)}(z,\overline{z}){\displaystyle \underset{r=1}{\overset{n}{}}}\left[r\mu _{\overline{z}}^{(r)}(z,\overline{z})C^{(nr+1)}(z,\overline{z})rC^{(r)}(z,\overline{z})\mu _{\overline{z}}^{(nr+1)}(z,\overline{z})\right]\right){\displaystyle \frac{\delta }{\delta \mu _{\overline{z}}^{(n)}(z,\overline{z})}}`$
$`+\left((\kappa _n^z+\overline{\kappa _n}^{\overline{z}}\overline{})\lambda _z^{Z^{(n)}}(z,\overline{z})+\lambda _z^{Z^{(n)}}(z,\overline{z})(\kappa _n^z+\mu (n,(z,\overline{z}))\overline{\kappa }^{\overline{z}})\right){\displaystyle \frac{\delta }{\delta \lambda _z^{Z^{(n)}}(z,\overline{z})}}`$
$`+\left[{\displaystyle \underset{\begin{array}{c}r,s=0\\ r+s1\end{array}}{\overset{n}{}}}c^{(r,s)}(z,\overline{z})(\tau _{n,(r,s)}^z+\overline{\tau }_{n,(r,s)}^{\overline{z}}\overline{})\phi (z,\overline{z})\right]{\displaystyle \frac{\delta }{\delta \phi (z,\overline{z})}}+c.c.]`$ (5.10)
(5.11)
where both the ghosts $`\kappa _n^z,๐^{(j)}`$ have been expressed in terms of the $`c^{(p,q)}`$ ghosts according to Eqs.(4.45)(4.26) respectively.
This is the ordinary diffeomorphism BRS operator, and its nilpotency is verified if the Beltrami conditions (3.28) hold for all the $`\lambda `$โs .
So the invariance of the Lagrangian $`\mathrm{\Gamma }_{Scalar}`$ in Eq (5.1) is written in a local form
$`\delta _c_{z,\overline{z}}(z,\overline{z})=(\kappa _n^z(z,\overline{z})_{z,\overline{z}}(z,\overline{z}))+\overline{}(\overline{\kappa }_n^{\overline{z}}(z,\overline{z})_{z,\overline{z}}(z,\overline{z}))`$ (5.12)
We now define a set of local operators of zero F-P charge by:
$`\delta _c={\displaystyle ๐z}d\overline{z}{\displaystyle \underset{\begin{array}{c}p,q=0\\ p+q1\end{array}}{\overset{n_{max}}{}}}\left(c^{(p,q)}(z,\overline{z})๐ฏ_{(p,q)}(z,\overline{z})+๐ฎc^{(p,q)}(z,\overline{z}){\displaystyle \frac{\delta }{\delta c^{(p,q)}(z,\overline{z})}}\right)`$ (5.15)
then thanks to both $`\{\delta _c,\delta _c\}=0`$, and Eq.(4.31), it turns out that the $`๐ฏ_{(p,q)}(z,\overline{z})`$โs fulfill commutation rules of $`w`$-algebra type, see e.g. and references therein :
$`[๐ฏ_{(p,q)}(z,\overline{z}),๐ฏ_{(r,s)}(z^{},\overline{z^{}})]=`$ (5.16)
$`=p_z^{}\delta ^{(2)}(z^{}z)๐ฏ_{(p+r1,q+s)}(z,\overline{z})r_z\delta ^{(2)}(zz^{})๐ฏ_{(p+r1,q+s)}(z^{},\overline{z^{}})`$
$`+q\overline{}_z^{}\delta ^{(2)}(z^{}z)๐ฏ_{(p+r,q+s1)}(z,\overline{z})s\overline{}_{\overline{z}}\delta ^{(2)}(zz^{})๐ฏ_{(p+r,q+s1)}(z^{},\overline{z^{}}).`$
Note however that the obtained $`w`$-algebra with respect to the $`c^{(p,q)}`$ ghosts is not chiral, but in the vacuum sector where the one relative to the $`C^{(j)}`$ ghosts is chiral.
###### Statement 5.2
The ordinary diffeomorphism BRS transformations will induce, from Eqs(4.9) (4.34) $`w`$-algebra symmetry transformations. The BRS functional operator to be used for the Field Theory quantum extension the diffeomorphism symmetry (in terms of the $`๐ฆ_n,\kappa _n`$ ghosts), will give, if written in terms of the $`c^{(r,s)}`$ ghosts, a BRS differential for $`w`$-algebras.
The ordinary procedure for Quantum extension suggests to introduce the anti-fields in the Lagrangian term:
$`_{antifields}={\displaystyle }dzd\overline{z}({\displaystyle \underset{r,s}{}}(\xi _{(r+1,s+1)}(z,\overline{z})๐ฎc^{(r,s)}(z,\overline{z}))+\nu _{(s+1)}(z,\overline{z})๐ฎ\mu _{\overline{z}}^{(s)}(z,\overline{z})`$
$`+\chi _{z,\overline{z}}(z,\overline{z})๐ฎ\phi (z,\overline{z})+{\displaystyle \underset{r}{}}(\rho (r,(z,\overline{z}))_{\overline{z},Z^{(r)}}๐ฎ\lambda _z^{(Z^{(r)})}(z,\overline{z}))+c.c.)`$ (5.17)
So the complete Classical Action becomes
$`\mathrm{\Gamma }^{Classical}=\mathrm{\Gamma }_{scalar}+\mathrm{\Gamma }_{antifields}`$ (5.18)
and at the classical level we get:
$`\delta \mathrm{\Gamma }^{(Classical)}{\displaystyle }dzd\overline{z}[{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \chi _{z,\overline{z}}(z,\overline{z})}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \varphi (z,\overline{z})}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \xi _{(r+1,s+1)}(z,\overline{z})}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta c^{(r,s)}(z,\overline{z})}}`$
$`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \nu _{(s+1)}(z,\overline{z})}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \mu _{\overline{z}}^{(s)}(z,\overline{z})}}+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \rho (r,(z,\overline{z}))_{\overline{z},Z^{(r)}}}}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \lambda _z^{Z^{(r)}}(z,\overline{z})}}+c.c.]=0`$
(5.19)
By the way if we want to reproduce here one of the outstanding feature of conformal models , that is the holomorphic properties of the object coupled in an invariant way to Beltrami fields (i.e. the Energy Momentum tensor) the task is not so simple.
This fact is, in this context, particularly fruitful : the presence of $`n`$ independent complex structures (and then $`n`$ independent Beltrami fields) means that we can derive at least $`n`$ energy-momentum tensors and their related holomorphic properties.
The problem is that the Beltrami multipliers are non local objects, so the Energy-Momentum tensor cannot be defined in terms of functional derivatives except for the case $`n=1`$.
We can provide a solution by the following shortcut : introduce the following lower triangular $`n_{max}\times n_{max}`$ matrix $`๐`$ with entries $`(r1,0)`$-differentials valued bilocal kernels - but highly non local in the $`\mu _{\overline{z}}^{(j)}`$โs,
$`๐(n,r;(z,\overline{z}),(z^{},\overline{z^{}}))_{(r1)}{\displaystyle \frac{\delta \mu _{\overline{z}}^z(n,(z,\overline{z}))}{\delta \mu _{\overline{z}}^{(r)}(z^{},\overline{z^{}})}},n=1\mathrm{}n_{max},0rn,`$ (5.20)
such that (compare with the expansion (3.24)),
$$\mu _{\overline{z}}^z(n,(z,\overline{z}))=๐z^{}d\overline{z}^{}\underset{r=1}{\overset{n}{}}๐(n,r;(z,\overline{z}),(z^{},\overline{z^{}}))_{(r1)}\mu _{\overline{z}}^{(r)}(z^{},\overline{z^{}}).\text{}$$
(5.21)
We shall suppose that $`๐`$ has an inverse $``$ with entries $`(2r,1)`$-differentials valued bilocal kernels, such that everywhere,
$`{\displaystyle ๐z\mathrm{"}}d\overline{z}\mathrm{"}{\displaystyle \underset{r=1}{\overset{n_{max}}{}}}(n,r;(z,\overline{z})(z\mathrm{"},\overline{z\mathrm{"}}))_{(2r,1)}๐(r,n^{};(z\mathrm{"},\overline{z\mathrm{"}}),(z^{},\overline{z^{}}))_{(r1,0)}=`$
$`=\delta _{n,n^{}}\delta ^{(2)}(zz^{}).`$ (5.22)
If we define:
$`๐ซ_z^{\overline{z}}(n,(z,\overline{z})){\displaystyle ๐z^{}}d\overline{z}^{}{\displaystyle \underset{r}{}}(n,r;(z,\overline{z})(z^{},\overline{z^{}}))_{(2r,1)}{\displaystyle \frac{\delta }{\delta \mu ^r(z^{},\overline{z^{}})}}`$ (5.23)
one thus has
$`๐ซ_z^{\overline{z}}(n,(z,\overline{z}))\mu (n^{}(z^{},\overline{z^{}}))=\delta _{n,n^{}}\delta ^{(2)}(zz^{}).`$ (5.24)
The $`๐ซ`$โs play the role of the โfunctional derivative operatorsโ with respect the Beltrami parameters; they will be (as well as these last) non local (in $`\mu ^{(r)}(z,\overline{z})`$ ) functional operators and have a fundamental role in our context. If $`\mu (n,(z,\overline{z}))`$is coupled at the tree approximation with a local field $`\mathrm{\Theta }_{(zz)}^{(Classical)}(n,(z,\overline{z}))`$ in an invariant way, for each n $`n=1\mathrm{}n_{max}`$ we have
$`\mathrm{\Theta }_{(zz)}^{(Classical)}(n,(z,\overline{z}))=๐ซ_z^{\overline{z}}(n,(z,\overline{z}))\mathrm{\Gamma }^{(Classical)}`$ (5.25)
so that the latter will transform at the classical level as:
$`๐ฎ\mathrm{\Theta }_{(zz)}^{(Classical)}(n,(z,\overline{z}))=๐ฆ^n(z,\overline{z})\mathrm{\Theta }_{zz}^{(Classical)}(z,\overline{z})+2\mathrm{\Theta }_{zz}^{(Classical)}(z,\overline{z})๐ฆ^n(z,\overline{z})`$ (5.26)
By the anticommutator $`\overline{}=\{๐ฎ,{\displaystyle \frac{}{\kappa _n}}\}`$ one gets :
$`\left(\overline{}\mu (n,(z,\overline{z}))2\mu (n,(z,\overline{z}))\right)\mathrm{\Theta }_{(zz)}^{(Classical)}(n,(z,\overline{z}))=0`$ (5.27)
Defining the $`(2,0)`$-differential
$`๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(n,(Z^{(n)},\overline{Z}^{(n)}))\left[{\displaystyle \frac{1}{\lambda _{z}^{Z^{(n)}}{}_{}{}^{2}}}\mathrm{\Theta }_{(zz)}^{(Classical)}(n,(z,\overline{z}))\right]`$ (5.28)
the previous equation leads to:
$`{\displaystyle \frac{}{\overline{Z}^n}}๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(n,(Z^{(n)},\overline{Z}^{(n)}))=0`$ (5.29)
In particular we remark that in the $`(z,\overline{z})`$ background:
$`d๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(Z^{(n)}(z,\overline{z}))=\left[{\displaystyle \frac{๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(Z^{(n)})}{Z^{(n)}}}\right](z,\overline{z})\lambda _z^{Z^{(n)}}(z,\overline{z})\left[dz+\mu (n,(z,\overline{z}))d\overline{z}\right]`$
(5.30)
This means that:
###### Statement 5.3
The conserved current $`๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}`$ is a non local function of $`\mu (n,(z,\overline{z}))`$. It will imply that this object is a nonlocal function of $`\mu _{\overline{z}}^{(j)}`$ for $`jn`$.
Moreover we can rewrite the B.R.S transformation of the current as:
$`๐ฎ๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(Z^{(n)})=\mathrm{{\rm Y}}^n_{Z^{(n)}}๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(Z^{(n)})`$ (5.31)
so that we can define a set of invariant charges:
$`๐ฌ_n^{(Classical)}={\displaystyle ๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(Z^{(n)})๐Z^{(n)}}`$ (5.32)
which are functional depending on the local parameters $`\mu _{\overline{z}}^{(j)}(z,\overline{z})`$ for $`jn`$,
$`๐ฎ๐ฌ_n^{(Classical)}=0,n=1\mathrm{}n_{max}`$ (5.33)
and a fortiori:
$`๐ฏ_{(p,q)}๐ฌ_n^{(Classical)}=0,p,qn=1\mathrm{}n_{max},`$ (5.34)
that is the charges $`๐ฌ_n`$ are invariant under both diffeomorphism and $`w`$ action.
Even if we have stressed the โnon localโ nature of our theory, we can ask whether some noteworthy property is hidden in the pure local sector of the model.
The local counterpart of the Energy-momentum tensor (which is invariantly coupled to the Beltrami fields) are the quantities which are coupled in an invariant way to the $`\mu _{\overline{z}}^{(j)}`$โs.
We have already pointed out that in this context all the geometrical architecture of our building cannot be appreciated;anyhow a relic of $`w`$ algebras still appears : we now show, that their OPEโs will generate a $`w`$ expansion,as it has already be shown in the Literature . Indeed introducing, at the Classical level, the Ward identity for the appropriate partition function $`๐ต^{(Classical)}(\mu )`$ of the vacuum sector is:
$`\overline{}{\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{(s)}(z,\overline{z})}}{\displaystyle \underset{j=0}{\overset{n_{max}s}{}}}\left((s+j+1)\mu _{\overline{z}}^{(j+1)}+(j+1)\mu _{\overline{z}}^{(j+1)}\right){\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{(j+s)}(z,\overline{z})}}=0`$
(5.35)
from which we derive by multiplying by $`\pi \frac{1}{(zz^{})}`$ and integration,
$`\pi ^2{\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{(s)}(z^{},\overline{z^{}})}}+{\displaystyle \underset{j=0}{\overset{n_{max}s}{}}}{\displaystyle ๐z}d\overline{z}\mu _{\overline{z}}^{(j+1)}(z,\overline{z})\left({\displaystyle \frac{s+j+1}{(zz^{})^2}}+{\displaystyle \frac{s}{(zz^{})}}\right)\pi {\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{z^{(j+s)}}(z,\overline{z})}}=0`$
Setting all the $`\mu _{\overline{z}}^{(j)}`$โs to zero and by quantum action principle one thus gets
$`\pi ^2{\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{(s)}(z^{},\overline{z^{}})\delta \mu _{\overline{z}}^{(r)}(z,\overline{z})}}|_{\mu =0}+{\displaystyle \underset{j=0}{\overset{n_{max}s}{}}}\delta _r^{j+1}\left({\displaystyle \frac{s+r}{(zz^{})^2}}+{\displaystyle \frac{s}{(zz^{})}}\right)\pi {\displaystyle \frac{\delta ๐ต^{Classical}(\mu )}{\delta \mu _{\overline{z}}^{(r+s1)}(z,\overline{z})}}|_{\mu =0}=0`$
which gives the OPE for the tensors coupled to these objects.
This is valid only at the classical level: the local theory display at the quantum level anomalies, while the โnon localโ approach admits, as we shall see in the next Section, a rather painless cancellation mechanism.
### 5.1 Quantum extension and Anomalies
The difficulties avoided using a local $`w`$ algebra using $`๐^{(n)}(z,\overline{z})`$ or $`c^{(r,s)}(z,\overline{z})`$ ghosts will create other problems for the Quantum extension of the model.
Due to quantum perturbative corrections the Action functional may violate the Ward identities, and according to the usual Lagrangian framework, one introduces the corresponding linearized BRS operator,
$`\delta \mathrm{\Gamma }=\mathrm{\Delta }`$ (5.38)
$`\delta {\displaystyle }dzd\overline{z}[{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \chi _{z,\overline{z}}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \varphi (z,\overline{z})}}+{\displaystyle \underset{r,s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \xi _{(r+1,s+1)}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta c^{(r,s)}(z,\overline{z})}}\right)`$
$`+{\displaystyle \underset{s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \nu _{(s+1)}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \mu _{\overline{z}}^{(s)}(z,\overline{z})}}\right)+{\displaystyle \underset{r}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \rho _{z,Z^{(r)}}(r,(z,\overline{z}))}}{\displaystyle \frac{\delta }{\delta \lambda _z^{Z^{(r)}}(z,\overline{z})}}\right)`$
$`+{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \varphi (z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \chi _{z,\overline{z}}(z,\overline{z})}}+{\displaystyle \underset{r,s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta c^{(r,s)}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \xi _{(r+1,s+1)}(z,\overline{z})}}\right)`$
$`+{\displaystyle \underset{s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \mu _{\overline{z}}^{(s)}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \nu _{(s+1)}(z,\overline{z})}}\right)+{\displaystyle \underset{r}{}}({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \lambda _z^{Z^{(r)}}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \rho _{z,Z^{(r)}}(r,(z,\overline{z}))}}+c.c.)]`$ (5.39)
so only by counter-terms inclusion the symmetry will be restored at each order of the perturbative expansion. As is well-known, this requires a cohomological approach and if the cohomology is empty, the symmetry is restored at the quantum level.
This calculation is performed in the Appendix, where we show that the cohomology sector in the space of the local functions is isomorphic to the tensor product of the cohomologies of the ordinary disjoint smooth changes of coordinates $`(z,\overline{z})(Z^{(r)}(z,\overline{z}),\overline{Z}^{(r)}(z,\overline{z})),r=1\mathrm{}n`$.
This result shifts the problem to the quantum extension of a theory whose symmetry is provided by $`n`$ disjoint ordinary changes of coordinates, for which many known results are at our disposal
In particular if we add to our field content all the $`\mathrm{ln}\lambda ^{Z(r)}(z,\overline{z})r=1\mathrm{}n_{max}`$ the cohomology becomes empty.
The implicitโnon locality on the fieldsโ of our model softens the possible disappointment coming from the introduction of logarithms.
In the usual Quantum Field Theory the local anomaly is a cocycle which has a coboundary term which is $`\mathrm{log}\lambda `$ dependent and derives from the usual transgression formulas coming from the Gelโfand-Fuchs cocycle, which becomes coboundary if we put $`\mathrm{ln}\lambda `$โs into the game: in our case we get:
$`\mathrm{\Delta }^{\mathrm{}}(z,\overline{z})`$ $`=`$ $`{\displaystyle \underset{r=1}{\overset{n}{}}}c_r๐ฆ_r(z,\overline{z})๐ฆ_r(z,\overline{z})^2๐ฆ_r(z,\overline{z})`$ (5.40)
$`=`$ $`{\displaystyle \underset{r=1}{\overset{n}{}}}c_r\delta \left(๐ฆ_r(z,\overline{z})๐ฆ_r(z,\overline{z})\mathrm{ln}\lambda _z^{(Z^r)}(z,\overline{z})\right)`$
modulo coboundary terms and total derivatives; the anomaly in the space of local functionals is recovered by using the techniques of Ref .
$`\mathrm{\Delta }(z,\overline{z})={\displaystyle \underset{r=1}{\overset{n_{max}}{}}}c_r๐ฆ_r^z(z,\overline{z})^3\mu _{\overline{z}}^z(r,(z,\overline{z}))`$
Anyhow the B.R.S diffeomorphism operator is deeply related to the one of $`w`$ symmetry when we render explicit the $`๐ฆ_n(z,\overline{z})`$ ghosts in terms of $`c^{(r,s)}(z,\overline{z})`$ ones; this allows to calculate the $`w`$ local anomalies:
$`๐ฏ_{(r,s)}(z,\overline{z})\mathrm{\Gamma }={\displaystyle \underset{n=\text{max}(r,s)}{\overset{n_{max}}{}}}c_n\eta _{(r,s)}(n,(z,\overline{z}))^3\mu (n,(z,\overline{z})).`$ (5.42)
From the usual construction
$`๐ฆ_r(z,\overline{z})๐ฆ_r(z,\overline{z})^2๐ฆ_r(z,\overline{z})=\delta (๐ฆ_r(z,\overline{z})๐ฆ_r(z,\overline{z})\mathrm{ln}\lambda _z^{Z^{(r)}}(z,\overline{z}))`$ (5.43)
so the non local Action compensating terms will be:
$`\mathrm{\Gamma }^{(Polyakov)}={\displaystyle }c_r{\displaystyle }dzd\overline{z}(\mu _{\overline{z}}^z(r,(z,\overline{z}))^2\mathrm{ln}\lambda ^{Z^{(r)}}(z,\overline{z})`$
and the corresponding e-m tensor
$`\mathrm{\Theta }_{(zz)}^{(Polyakov)}(n,(z,\overline{z}))=๐ซ_z^{\overline{z}}(n,(z,\overline{z}))\mathrm{\Gamma }^{(Polyakov)}=2c_nS_{zz}(Z^{(n)}(z,\overline{z}))`$ (5.45)
where, as usual:
$`S_{zz}(Z^{(n)}(z,\overline{z}))^2\mathrm{ln}\lambda _z^{Z^{(n)}}(z,\overline{z}){\displaystyle \frac{1}{2}}\left(\mathrm{ln}\lambda _z^{Z^{(n)}}(z,\overline{z})\right)^2`$ (5.46)
So we can define:
$`\mathrm{\Theta }_{(zz)}(n,(z,\overline{z}))=๐ซ_z^{\overline{z}}(n,(z,\overline{z}))\left[\mathrm{\Gamma }\mathrm{\Gamma }^{(Polyakov)}\right]`$ (5.47)
So the anomaly compensation mechanism allows to construct at the Quantum level an energy momentum tensor which verifies the same symmetry properties as the classical energy momentum tensor.
At this stage it is trivial to define a current
$`๐ฅ_{Z^{(n)}Z^{(n)}}(n,(Z^{(n)},\overline{Z}^{(n)})){\displaystyle \frac{1}{\lambda _{z}^{Z^{(n)}}{}_{}{}^{2}}}\left[๐ซ_z^{\overline{z}}(n,(z,\overline{z}))\mathrm{\Gamma }+2c_nS_{zz}(Z^{(n)}(z,\overline{z}))\right]`$ (5.48)
which is the Quantum extension of the classical $`(2,0)`$-covariant tensor $`๐ฅ_{Z^{(n)}Z^{(n)}}^{(Classical)}(n,(Z^{(n)},\overline{Z}^{(n)}))`$. Note that due to the presence of the Schwarzian derivative the former is no longer a tensor.
Moreover we can also define, for each $`nn_{max}`$:
$`๐ฌ_n={\displaystyle ๐ฅ_{Z^{(n)}Z^{(n)}}(Z^{(n)})๐Z^{(n)}}`$ (5.49)
which will be invariant even in the Quantum level.
## 6 Conclusions
The many aspects of two dimensional reparametrization invariance provide a further geometrical description of $`w`$-algebras. We have addressed the question of introducing local $`(n,1)`$-conformal fields generalizing the usual Beltrami differential appearing in $`w`$-gravity. It was shown that the way out is based on the infinitesimal action of symplectomorphisms on coordinate transformations dictated by very special canonical transformations.
Also, it is both interesting and intriguing to note how intermingled the symplectic and conformal geometries are relevant for all the present treatment. The combination of Beltrami parametrization of complex structures, canonical transformations and symplectomorphisms yields to a BRS formulation of $`w`$-algebras.
However, although the locality requirements are fundamental for the physical contents within a Lagrangian field Theory, we have overcome them in order to take ever present the geometrical aspect of the problem. But we aim to treat the former in order to understand better the role of the Quantum $`w`$ local anomalies in relation to the point of view expressed in the present paper.
Acknowledgements. We are grateful to Prof. A. Blasi for comments and discussions.
## 7 Appendix
The purpose of this Appendix is to show that the cohomology space of our BRS operator $`\delta `$ in the space of local functions, is isomorphic to the one of the $`n`$ independent reparametrizations $`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)})`$.
We have shown in that the cohomology space in the functional of the BRS operator $`\delta `$ will coincide with the local function cohomology of the nilpotent BRS operator $`\delta c^{(1,0)}c^{(0,1)}\overline{}`$.
This cohomology space will be computed by using the spectral sequences method. Let us filter with:
$`\nu ={\displaystyle \underset{p,q,m,n}{}}\left(p+q\right)^m\overline{}^nc^{(p,q)}(z,\overline{z}){\displaystyle \frac{}{^m\overline{}^nc^{(p,q)}(z,\overline{z})}}.`$ (7.1)
At the zero eigenvalue the following operator is obtained,
$`\delta _0{\displaystyle }dzd\overline{z}[{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \varphi (z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \chi _{z,\overline{z}}(z,\overline{z})}}+{\displaystyle \underset{s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \mu _{\overline{z}}^{(s)}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \nu _{(s+1)}(z,\overline{z})}}\right)`$
$`+{\displaystyle \underset{r,s}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Antifields)}}{\delta c^{(r,s)}(z,\overline{z})}}|_{c=0}{\displaystyle \frac{\delta }{\delta \zeta _{(r+1,s+1)}(z,\overline{z})}}\right)+{\displaystyle \underset{r}{}}\left({\displaystyle \frac{\delta \mathrm{\Gamma }^{(Classical)}}{\delta \lambda _z^{Z^{(r)}}(z,\overline{z})}}{\displaystyle \frac{\delta }{\delta \rho _{z,Z^{(r)}}(r,(z,\overline{z}))}}\right)]`$ (7.2)
where $`{\displaystyle \frac{\delta \mathrm{\Gamma }^{(Antifields)}}{\delta c^{(p,q)}(z,\overline{z})}}|_{c=0}`$ is the $`c`$ independent part of the BRS variation of the anti-fields $`\zeta `$ induced by the linearization of $`\delta `$.
This operator is clearly nilpotent due to the $`\mathrm{\Phi }`$-$`\mathrm{\Pi }`$ neutrality of the $`\mathrm{\Gamma }^{Classical}`$ terms. Its cohomology space can be calculated using again the spectral sequences method. Its adjoint can be defined upon using the Dixon procedure and the Laplacian kernel is isomorphic to the cohomology space. The upshot of this calculation does not modify the final result; anyhow for the sake of completeness we can calculate this space by first filtrating this operator with the field operator counter, and by calculating the kernel of the Laplacian. It is easy to convince one self that the cohomology space will be independent on the anti-fields $`\rho _{z,Z^{(r)}}(r,(z,\overline{z}))`$,$`\nu _{(s+1)}(z,\overline{z})`$, $`\chi _{z,\overline{z}}(z,\overline{z})`$, $`\zeta _{r+1,s+1}(z,\overline{z})`$ and complicated combinations in the matter fields and $`\lambda `$โs and $`\mu `$โs.
The fundamental step takes place in the analysis of the action on $`\delta `$ of the filtering operator (7.1) at the eigenvalue equal to one. In this case we have to calculate the kernel of this operator (and its adjoint) on the space previously calculated with $`\delta _0`$ . It is easy to derive that this operator is nothing else but the sum of the operators $`\delta c^{(1,0)}c^{(0,1)}\overline{}`$, (where $`\delta `$ is the ordinary diffeomorphism operator of the $`\mathrm{\Phi }`$-$`\mathrm{\Pi }`$ neutral fields containing the ghosts $`c^{(1,0)}`$ and $`c^{(0,1)}`$) plus the total variation of $`c^{(p,q)}`$.
This operator is still nilpotent so we can filter it again. We shall choose as filtering operator the one which counts the $`c^{(1,0)}`$ and $`c^{(0,1)}`$ ghost fields, namely,
$`\nu ^{}={\displaystyle \underset{p,q,m,n}{}}^n\overline{}^mc^{(1,0)}(z,\overline{z}){\displaystyle \frac{}{^n\overline{}^mc^{(1,0)}(z,\overline{z})}}+^n\overline{}^mc^{(0,1)}(z,\overline{z}){\displaystyle \frac{}{^n\overline{}^mc^{(0,1)}(z,\overline{z})}}`$ (7.3)
At zero eigenvalue we find:
$`\delta _0^{}={\displaystyle \underset{\begin{array}{c}j,l,m,n,r,s,p,q\\ p+q>r+s>1\end{array}}{}}\frac{n!m!}{l!j!(nl)!(mj)!}[^l\overline{}^jc^{(r,s)}(z,\overline{z})(r^{(nl+1)}\overline{}^{(mj)}c^{(pr+1,qs)}(z,\overline{z})`$ (7.6)
$`+s^{(nl)}\overline{}^{(mj+1)}c^{(pr,qs+1)}(z,\overline{z}))]{\displaystyle \frac{}{(^n\overline{}^mc^{(p,q)}(z,\overline{z}))}}`$ (7.7)
After defining its adjoint according to the Dixon procedure it is easy to find that the cohomology does not depend on the ghost fields $`c^{(p,q)}`$ and their derivatives, with the condition $`\left(p+q\right)>1`$.
At the end we are left with the BRS operator induced by the following transformation rules, for any $`n`$:
$`๐ฎ\mu _{\overline{z}}^{(n)}(z,\overline{z})=c^{(1,0)}(z,\overline{z})\mu _{\overline{z}}^{(n)}(z,\overline{z})+\overline{}\left(\mu _{\overline{z}}^{(n)}(z,\overline{z})c^{(0,1)}(z,\overline{z})\right)+\overline{}c^{(1,0)}(z,\overline{z})\delta _{n,1}`$
$`n\mu _{\overline{z}}^{(n)}(z,\overline{z})c^{(1,0)}(z,\overline{z})\left({\displaystyle \underset{r}{}}r\mu _{\overline{z}}^{(r)}(z,\overline{z})\mu _{\overline{z}}^{(nr+1)}(z,\overline{z})\right)c^{(0,1)}(z,\overline{z})`$
(7.8)
$`๐ฎ\lambda _z^{Z^{(n)}}(z,\overline{z})`$ $`=`$ $`\left(c^{(1,0)}(z,\overline{z})+c^{(0,1)}(z,\overline{z})\overline{}\right)\lambda _z^{Z^{(n)}}(z,\overline{z})`$ (7.9)
$`+\lambda _z^{Z^{(n)}}(z,\overline{z})\left(c^{(1,0)}(z,\overline{z})+\mu (n,(z,\overline{z}))c^{(0,1)}(z,\overline{z})\right)`$
where $`\mu (n,(z,\overline{z}))`$ must be written according to the expansion Eq.(3.24), we thus get the transformations for the Beltrami differentials at any level $`n`$:
$`๐ฎ\mu (n,(z,\overline{z}))=\left(c^{(1,0)}(z,\overline{z})+c^{(0,1)}(z,\overline{z})\overline{}\right)\mu (n,(z,\overline{z}))`$
$`+\overline{}c^{(1,0)}(z,\overline{z})+\mu (n,(z,\overline{z}))\overline{}c^{(0,1)}(z,\overline{z})`$
$`\mu (n,(z,\overline{z}))\left(c^{(1,0)}(z,\overline{z})+\mu (n,(z,\overline{z}))c^{(0,1)}(z,\overline{z})\right)`$ (7.10)
while for the scalar matter field,
$`๐ฎ\phi (z,\overline{z})=\left(c^{(1,0)}(z,\overline{z})+c^{(0,1)}(z,\overline{z})\overline{}\right)\phi (z,\overline{z})`$ (7.11)
and the ghost field $`c^{(1,0)}`$ (and the c.c. for $`c^{(0,1)}`$),
$`๐ฎc^{(1,0)}(z,\overline{z})=\left(c^{(1,0)}(z,\overline{z})_z+c^{(0,1)}(z,\overline{z})\overline{}_{\overline{z}}\right)c^{(1,0)}(z,\overline{z}).`$ (7.12)
All of these are ordinary diffeomorphism transformations.
So the $`w`$ algebra reduces to a tensor product of $`n`$ independent diffeomorphisms of level equal to one $`(z,\overline{z})(Z^{(n)},\overline{Z}^{(n)}),n=1,\mathrm{},n_{max}`$. |
warning/0003/cond-mat0003028.html | ar5iv | text | # On the dynamics of coupled S =1/2 antiferromagnetic zig-zag chains
## I Introduction
Spin systems consisting of chain- or ladderlike structures as basic building blocks have recently attracted much attention. These systems are of interest on the one hand as one-dimensional (1D) model systems allowing to study quantum phase transitions related to the existence of a spin gap and their dependence on the exchange parameters ; on the other hand they describe an increasing number of real materials when an additional (small) exchange coupling in the remaining two dimensions is introduced . A material of particular recent experimental interest is $`\mathrm{KCuCl}_3`$ .
We have performed an investigation of the dynamics of such systems with a twofold aim: (i) We discuss the spectrum $`\omega (q_x)`$ of the low-lying triplet excitations in the ideal 1D system over a wide range of exchange parameters using both series expansions and exact diagonalization, in order to determine the range of applicability of the series expansion approach and to study the validity of using an effective interaction between dimers. We find and discuss in particular a regime in phase space with extremely small dispersion and a minimum of $`\omega (q_x)`$ at finite wavevector $`0<q_x<\pi `$. (ii) We extend the dimer expansion to include 3D couplings and apply this method in particular to a discussion of the dynamics of the quasi zigzag-ladder material KCuCl<sub>3</sub> in terms of microscopic exchange parameters.
Of particular interest as an 1D building block for this type of materials is the $`S=\frac{1}{2}`$ zig-zag chain, as shown in Fig. 1 and defined by the following hamiltonian:
$`H={\displaystyle \underset{i=1}{\overset{L}{}}}J\stackrel{}{S}_{1,i}\stackrel{}{S}_{2,i}`$ $`+`$ $`J_1\stackrel{}{S}_{1,i}\stackrel{}{S}_{2,i+1}`$ (1)
$`+`$ $`J_2\left(\stackrel{}{S}_{1,i}\stackrel{}{S}_{1,i+1}+\stackrel{}{S}_{2,i}\stackrel{}{S}_{2,i+1}\right).`$ (2)
On the theoretical side, this generic 1D model interpolates between a number of seemingly different limiting cases: It is an alternative way to formulate the hamiltonian for the generalized $`S=\frac{1}{2}`$ spin ladder generalized to include one diagonal interaction or equivalently the $`S=\frac{1}{2}`$ chain with nearest neighbour (NN) alternating exchange and next nearest neighbour exchange (we use the shorthand NNNA-chain in the following). It thus covers the well-known limiting models of the isotropic $`S=\frac{1}{2}`$ Heisenberg chain (HAF, $`J=J_1,J_2=0`$), the standard antiferromagnetic $`S=\frac{1}{2}`$ ladder ($`J_1=0,J=J_2>0`$), the weakly interacting dimer chain ($`J_1,J_2J`$) and the $`S=1`$ antiferromagnetic (Haldane) chain ($`J_1\mathrm{},J+2J_2>0`$). It can alternatively be considered as a two legged spin ladder with rung coupling $`J`$, leg coupling $`J_2`$ and additional diagonal coupling $`J_1`$.
The theoretical interest in the dynamics of the NNNA-chain goes back to the work of Shastry and Sutherland , who identified the elementary excitations without alternation as free particles (spinons), which may become bound. A variational approach to the excitations of the NNNA-chain based on this concept has recently been shown to cover qualitatively the transition from free spinons to the Haldane triplet.
In a first approximation real materials are often considered as examples of 1D chains with a hamiltonian as given in Eq. (1); then they realize different points in the phase diagram spanned by the interaction constants $`J_1,J_2`$ and illustrate the interest to describe systematically the variation of static and dynamic properties with the parameters $`J_1,J_2`$. Static properties, such as susceptibility and specific heat, however, have turned out to be rather insensitive to the details of the microscopic hamiltonian and, in the case of $`(\mathrm{VO})_2\mathrm{P}_2\mathrm{O}_7`$ have even not been able to reveal the basic interactions as two- instead of one-dimensional.
Thus for a description of real materials a systematic microscopic treatment of the dynamics is of particular importance. In section II we present a systematic overview of the dynamics of the 1D system, discussing both general properties as well as comparing results from exact diagonalization to results from series expansions. In section III we use the series expansion approach to calculate the low-lying excitations in a 3D material with the structure of $`\mathrm{KCuCl}_3`$ in terms of the microscopic hamiltonian. This will allow us to go beyond the determination of effective dimer exchange parameters in recent work and to determine the microscopic exchange parameters. A summary will be given in the concluding section IV.
## II Elementary excitations of the 1D zig-zag chain
We start with a short summary of the symmetries of the 1D system (the chain direction is denoted as $`x`$axis): Translational symmetry is described by the wavevector $`q_x`$ defined in a Brillouin zone $`\frac{\pi }{a}<q_x<+\frac{\pi }{a}`$; the unit cell of length $`a`$ contains two spins, or equivalently one dimer (one singlet in the limit $`J_1=J_2=0`$). We thus use the conventional notation for ladders: $`a`$ is the distance between rungs, whereas the distance between spins in the NNNA-chain picture is $`a/2`$. We thus expect two basic excitations per unit cell. We will use units $`a=1`$ in the following. Excitation frequencies at wavevectors $`q_x`$ and $`q_x`$ are equal owing to reflection symmetry along the chain.
For special points in the phase diagram additional symmetries exist:
Without alternation ($`J_1=J`$) it is natural to use a unit cell of length $`\stackrel{~}{a}=\frac{1}{2}a`$ containing only one spin. Our Brillouin zone is half of this Brillouin zone of the uniform chain and the excitations of the conventional spin chain will appear folded back to our smaller Brillouin zone.
For the ladder symmetry ($`J_1=0`$) there exists a quantum number parity, $`P`$, resulting from the interchange of the two legs and we can classify states as positive or negative under this reflection. Each dimer in the singlet (triplet) state contributes a factor of $`1(+1)`$ to this parity. An alternative notation introduces the component $`q_{}`$ with values 0 (corresponding to $`P=+1`$) and $`\pi `$ (corresponding to $`P=1`$).
The ground state is a singlet in the whole phase plane and the lowest excited state is generally a triplet. The ground states for the ladder symmetry ($`J_1=0`$) have parity $`P=+1`$ for $`L`$ even.
We have studied the dispersion $`\omega (q_x)`$ of the basic triplet excitation for a typical variety of paths in the $`J_1J_2`$ parameter space by two methods: (i) By exact numerical diagonalization, using the Lanczos algorithm, we have calculated $`\omega (q_x)`$ for the lowest excited states (between 2 and 4 states) for 24 spins, i.e. for 7 different values of wavevector $`q_x`$. (ii) We have performed series expansions around the dimer point, $`J_1=J_2=0`$, up to third order analytically and up to 10th order after implementation of the cluster algorithm on an Alpha work station. Thus we have obtained the ground state energy $`E_0`$ and an effective Hamiltonian which can be diagonalized by a Fourier transformation. Finally we get the dispersion relation for the lowest excited state expressed as series $`_na_n\mathrm{cos}(nq_x)`$.
In the follwoing we present a number of results for the 1D zig-zag chain which prepare the stage for the first application of the method to a nontrivial 3D system in section III and also add some new aspects to the large number of previous studies on the 1D system defined by eq. (1) in recent years. To give a short review of existing work we mention first that the dimer series expansion approach started when the work of Brooks Harris was revived by Uhrig in the context of $`\mathrm{CuGeO}_3`$. The expansion for the triplet dispersion was extended to high orders recently by Oitmaa et al for ladders to $`8^{\mathrm{th}}`$ order, by Barnes et al for the Heisenberg alternating chain to $`9^{\mathrm{th}}`$ order and by Singh et al for the disorder line to $`23^{\mathrm{th}}`$ order (using a special symmetry on this line). The model of Eq. (1) was also treated by alternative methods as random phase approximation , Brรผckner theory for the equivalent dilute Bose gas , exact diagonalization , and continued fraction expansion based on ED results and DMRG . From these studies a rather complete picture of the low-energy dynamics of the 1D zigzag chain has emerged. In this section we supplement this picture by the following two remarks.
We start from the neighbourhood of the dimer point where an expansion in $`J_1,J_2`$ to low orders is sufficient. Up to third order the following result for the dispersion is obtained (as given in ref. , frequency and exchange constants are measured in units of the intradimer exchange $`J`$ from now on):
$`\omega (q_x)`$ $`=`$ $`1{\displaystyle \frac{J_1^2}{4}}\left(1+J_2\right)+{\displaystyle \frac{3}{8}}\left(J_2{\displaystyle \frac{J_1}{2}}\right)^2\left(2+J_2{\displaystyle \frac{J_1}{2}}\right)`$ (3)
$`+`$ $`\left[J_2{\displaystyle \frac{J_1}{2}}{\displaystyle \frac{J_1^2}{4}}\left(1+J_2\right){\displaystyle \frac{1}{4}}\left(J_2{\displaystyle \frac{J_1}{2}}\right)^3\right]\mathrm{cos}q_x`$ (4)
$``$ $`{\displaystyle \frac{1}{4}}\left(J_2{\displaystyle \frac{J_1}{2}}\right)^2\left(1+J_2+{\displaystyle \frac{J_1}{2}}\right)\mathrm{cos}2q_x`$ (5)
$`+`$ $`{\displaystyle \frac{1}{8}}\left(J_2{\displaystyle \frac{J_1}{2}}\right)^3\mathrm{cos}3q_x.`$ (6)
In first order perturbation theory in $`J_1`$ and $`J_2`$ the spectrum is dispersionless on the Shastry-Sutherland line $`J_1=2J_2`$ (also known as โdisorder lineโ). Eq. (2) also shows the general feature that the location of the minimum of the dispersion curve shifts from $`q_x=0`$ to $`q_x=\pi `$ somewhere close to crossing this line.
In recent approximate theoretical treatments of interacting dimer systems the exchange interactions between individual were reduced to an effective interaction between dimers. For the present 1D zigzag chain the result of this approximation is that only the combination $`J_2\frac{1}{2}J_1`$ enters into the dispersion. In more detail, the result is
$$\omega (q_x)=\sqrt{1+2\delta \omega ^{(1)}(q_x)}$$
(7)
where $`1+\delta \omega ^{(1)}(q_x)`$ is the dispersion in lowest order, i.e. the result of simple propagation of an excited dimer triplet without considering its coupling to higher energy modes. Evidently this is true in lowest order of the expansion; it is seen, however, already from Eq. (2) that additional terms which depend on the individual exchange interactions enter in higher order. Comparing the higher order coefficients in the series expansion we find that the effective dimer approximation of Eq. (3) amounts to keeping only the leading (i.e. lowest) powers in $`J_1,J_2`$ for each coefficient $`a_n`$ of $`\mathrm{cos}(nq_x)`$.
For a quantitative check of the effective dimer approximation we refer first to ref. where the dispersion on the disorder line, i.e. for fixed $`J_2\frac{1}{2}J_1=0`$ is shown. The effective dimer approximation is reasonable for a large part of the line but deteriorates rapidly when the non-alternating limit $`J_1=1`$, i.e. the Majumdar-Ghosh point is approached. For the more generic value $`J_2\frac{1}{2}J_1=\frac{3}{8}`$ we show the dispersion in Fig. 2; it is seen that the effective dimer approximation is of very limited value in this case. In addition these results, comparing the series expansion results to spectra from exact diagonalization, we find the following limits of validity for the dimer series expansion method: Whereas the $`10^{\mathrm{th}}`$ order spectra coincide with exact diagonalization results within a few percent for $`|J_1|1`$ and $`J_2<0.5`$ the accuracy deteriorates rapidly when the symmetric ladder ladder configuration ($`J_1=0,J_2=1`$) is approached.
Whereas the dispersion is flat on the disorder line to first order in $`J_1,J_2`$, but develops maxima and minima in higher orders with an apparent jump of the minimum energy form $`q_x=0`$ to $`q_x=\pi `$ close to this line, it was noted already before from exact diagonalization results that a regime with extremely flat dispersion and a minimum at finite wavevector $`0<q_x<\pi `$ might exist in a narrow regime in $`J_1J_2`$ parameter space. We have investigated this point again using the series expansions to $`10^{\mathrm{th}}`$ order which provide us with a continuous wavevector dependence and have confirmed the earlier speculation: In Fig. 3, 4 we show two examples for spectra on the lines $`J_1=12J_2`$ resp. $`J_1=\frac{1}{2}(1J_2)`$: The points in Fig. 3 were discussed before in ref. ; Fig. 3 demonstrates that the shallow minimum of the dispersion curve at a wavevector in the middle of the Brillouin zone is reproduced in the series expansions. As further example we show in Fig. 4 the flat dispersion for somewhat smaller values of $`J_1`$. In Fig. 5 we show the development of the wavevector $`q_{\mathrm{min}}`$ for the minimum value of $`\omega (q)`$ on the line $`J_1=\frac{1}{2}(1J_2)`$ in the $`J_1J_2`$ parameter space (corresponding to the case of Fig. 4). The gradual transition of $`q_{\mathrm{min}}`$ between $`q_{\mathrm{min}}=0`$ and $`q_{\mathrm{min}}=\pi `$, possibly with infinite slope at the end points, is clearly seen. The comparison between results in $`8^{\mathrm{th}}`$ and $`10^{\mathrm{th}}`$ order illustrate the convergence of the expansion. At present, however, the series expansions do not give a hint to a possible fundamental reason for nearly flat dispersion: We have examined the expansion coefficients $`a_n`$ of $`\mathrm{cos}(nq_x)`$ up to $`10^{\mathrm{th}}`$ order, but we do not find any indication of a convergence to zero for $`n0`$.
## III Interacting zig-zag Chains
In real materials consisting of weakly interacting chains such as $`\mathrm{KCuCl}_3`$ and $`\mathrm{CuGeO}_3`$ it is clear from inelastic neutron scattering experiments that there is considerable dispersion for wavevectors perpendicular to the double chain direction. For weakly interacting chains, the series expansion approach is to be considered as the only reliable systematic approach which (by comparing results in subsequent orders) allows a consistency check. Whereas the series expansions have been extended to cover systems coupled in 2D and interesting results have been obtained for the spin Peierls material $`\mathrm{CuGeO}_3`$ , for the 1/5 depleted square material $`\mathrm{CaV}_4\mathrm{O}_9`$ as well as for general parameters , we present in the following the first results using the expansion for weakly interacting dimers for a 3D coupled system with particular application to investigate the magnon dispersion in the material $`\mathrm{KCuCl}_3`$. The zig-zag chain system $`\mathrm{KCuCl}_3`$ actually appears to be closer to the dimer point than the systems treated so far and therefore is supposed to be a better candidate for this expansion.
The structure of $`\mathrm{KCuCl}_3`$ is shown schematically in Fig. 6 in a projection to what is conventionally called the $`xz`$plane: The fundamental dimers which are shown as solid lines form zig-zag chains in the $`x`$direction, neighbouring zig-zag chains are shifted with respect to each other in $`y`$direction by half a lattice constant $`\frac{b}{2}`$; this shift as well as the tilting of the internal dimer direction are indicated in Fig. 6 by giving the $`y`$coordinate for each line of spins in $`x`$direction ($`n`$ is an integer which numbers the different planes). The elementary cell consists of two dimers, dimer $`D_1`$ at the origin $`\stackrel{}{R}_1=0`$ and dimer $`D_2`$ at position $`\stackrel{}{R}_2=(0,\frac{1}{2},\frac{1}{2})`$ and the two spins forming each dimer, 1 and 2, are at positions $`\stackrel{}{R}_i+\stackrel{}{d}_i`$ for spin 1 and $`\stackrel{}{R}_i\stackrel{}{d}_i`$ for spin 2, for definiteness we take $`d_{i,z}>0`$.
For the exchange interactions between spins we will use the following notation: The main intradimer exchange is denoted as $`J`$. The exchange interaction per bond between spins in dimers separated by a lattice vector $`\stackrel{}{R}=la\stackrel{}{e}_x+mb\stackrel{}{e}_y+nc\stackrel{}{e}_z`$ is denoted as $`J_{(lmn)}`$ for the exchange between equivalent spins (pairs (11) or (22)) of the corresponding dimers and as $`J_{(lmn)}^{}`$ for the exchange between nonequivalent spins (pairs (12) or (21)). The following exchange interactions will be considered: $`J_{(100)}`$ for pairs (11) and (22), $`J_{(100)}^{}`$ for the pair (12) and $`J_{(201)}^{}`$ for the pair (12), and for the two cases $`p=0`$ and $`p=1`$: $`J_{(p\frac{1}{2}\frac{1}{2})}`$ for the pair (11) starting from dimer $`D_1`$ and for the pair (22) starting form dimer $`D_2`$ , $`J_{(p\frac{1}{2}\frac{1}{2})}=J_{(p\frac{1}{2}\frac{1}{2})}`$ for the pair (22) starting from dimer $`D_1`$ and for the pair (11) starting from dimer $`D_2`$, $`J_{(p\frac{1}{2}\frac{1}{2})}^{}=J_{(p\frac{1}{2}\frac{1}{2})}^{}`$ for pairs (12) starting from either dimer $`D_1`$ or dimer $`D_2`$.
Alternatively we can look at the structure projecting on the plane spanned by the directions $`\stackrel{}{e}_y`$ and $`\stackrel{}{e}_x+\frac{1}{2}\stackrel{}{e}_z`$. In a schematic picture which shows the topology of the exchange interactions only, we obtain Fig. 7; the 3D structure of $`\mathrm{KCuCl}_3`$ results when identical planes are stacked and the fundamental dimers are connected by zig-zag interactions. Evidently the planar structure of Fig. 7 can be reduced to a number of limiting cases including coupled alternating chains ($`J_{(1\pm \frac{1}{2}\frac{1}{2})}=0`$) or coupled ladders ($`J_{(201)}=J_{(1\frac{1}{2}\frac{1}{2})}^{}=0`$).
If the dispersion is considered only to first order or if higher orders are included in an RPA-like approximation , $`J`$ and $`J^{}`$ for a given $`(lmn)`$ enter only in the combination $`J^{\mathrm{eff}}=\frac{1}{2}(pJp^{}J^{})`$, where $`p,p^{}=1,2`$ is the number of exchange paths between the interacting dimers; $`J^{\mathrm{eff}}`$ is denoted as effective dimer interaction. As already seen in the 1D case of section 2, a correct treatment beyond first order involves $`J`$ and $`J^{}`$ independently. According to previous work the main exchange interactions in addition to the basic intradimer exchange are between dimers separated by $`(lmn)=(100),(201),(1\pm \frac{1}{2}\frac{1}{2})`$. Except for $`(201)`$ these dimer-dimer interactions involve both $`J`$ and $`J^{}`$ and it is our aim in the following to discuss the validity of the effective dimer approximation for $`\mathrm{KCuCl}_3`$ and to investigate to what extent the exchange interactions between individual spins can be determined from the present status of experimental results.
A calculation up to second order in the ratios $`J_{(lmn)}/J`$ leads to the following expression for the frequency of the basic triplet (frequency and coupling constants are measured in units of the basic dimer exchange constant $`J`$ and wavevectors $`q_i`$ are given in units with the crystallographic lattice constants $`a,b,c`$ set equal to unity):
$`\omega (\stackrel{}{q})`$ $`=1+\delta \omega ^{(1)}(\stackrel{}{q}){\displaystyle \frac{1}{2}}\delta \omega _{}^{(1)}{}_{}{}^{2}(\stackrel{}{q})`$ (8)
$`+`$ $`J_{(100)}\left(J_{(100)}J_{(100)}^{}\right){\displaystyle \frac{1}{4}}J_{}^{}{}_{(100)}{}^{2}\mathrm{cos}q_x`$ (9)
$`+`$ $`J_{\left(0\frac{1}{2}\frac{1}{2}\right)}J_{\left(0\frac{1}{2}\frac{1}{2}\right)}^{}+J_{\left(1\frac{1}{2}\frac{1}{2}\right)}J_{\left(1\frac{1}{2}\frac{1}{2}\right)}^{}`$ (10)
$`+`$ $`{\displaystyle \frac{1}{2}}\left(J_{\left(0\frac{1}{2}\frac{1}{2}\right)}^2J_{}^{}{}_{\left(0\frac{1}{2}\frac{1}{2}\right)}{}^{2}\right)\mathrm{cos}{\displaystyle \frac{q_y}{2}}\mathrm{cos}{\displaystyle \frac{q_z}{2}}`$ (11)
$`+`$ $`{\displaystyle \frac{1}{2}}\left(J_{\left(1\frac{1}{2}\frac{1}{2}\right)}^2J_{}^{}{}_{\left(1\frac{1}{2}\frac{1}{2}\right)}{}^{2}\right)\mathrm{cos}{\displaystyle \frac{q_y}{2}}\mathrm{cos}{\displaystyle \frac{2q_x+q_z}{2}}`$ (12)
$``$ $`{\displaystyle \frac{1}{4}}J_{}^{}{}_{(201)}{}^{2}\mathrm{cos}(2q_x+q_z)`$ (13)
Here
$`\delta \omega ^{(1)}(\stackrel{}{q})`$ $`={\displaystyle \frac{1}{2}}\left(2J_{(100)}J_{(100)}^{}\right)\mathrm{cos}q_x`$ (14)
$`+`$ $`\left(J_{\left(0\frac{1}{2}\frac{1}{2}\right)}J_{\left(0\frac{1}{2}\frac{1}{2}\right)}^{}\right)\mathrm{cos}{\displaystyle \frac{q_y}{2}}\mathrm{cos}{\displaystyle \frac{q_z}{2}}`$ (15)
$`+`$ $`\left(J_{\left(1\frac{1}{2}\frac{1}{2}\right)}J_{\left(1\frac{1}{2}\frac{1}{2}\right)}^{}\right)\mathrm{cos}{\displaystyle \frac{q_y}{2}}\mathrm{cos}{\displaystyle \frac{2q_x+q_z}{2}}`$ (16)
$``$ $`{\displaystyle \frac{1}{2}}J_{\left(201\right)}^{}\mathrm{cos}\left(2q_x+q_z\right)`$ (17)
is the dispersion to first order in the exchange constants, i. e. the effect of simple propagation of an excited dimer triplet.
In Eq. 14 we have used for simplicity an extended zone scheme: Since there are two dimers in the proper crystallographic unit cell, there are two branches of triplet excitations for each wave vector in the crystallographic Brillouin zone ($`\pi <q_yb,q_zc\pi `$). Owing to the symmetry of the exchange interactions in the hamiltonian, these two branches join to the unique smooth expression given in Eq. 14, when the doubled zone $`2\pi <q_zc+q_yb,q_zcq_yb2\pi `$ is used. Each triplet excitation, however, is present at all equivalent wavevectors of the crystallographic reciprocal lattice. For a discussion of dipole transition amplitudes the noninteracting triplet approximation can be used as sufficient guide and gives the follwing results : For momentum tranfer perpendicular (parallel) to $`\stackrel{}{e}_y`$ the branch of Eq. 14 in the first (second) crystallographic Brillouin zone gives the only nonvanishing contribution. (A change form the first to second crystallographic Brillouin zone then corresponds to a change in either $`q_y`$ or $`q_z`$ by $`2\pi `$, i.e. to a change of sign in the second and third line of Eq. 14).
We note that the wavevector $`q_z`$ should be distinguished from the quantity $`q_{}`$ which is often used to denote the two values of the quantum number parity discussed in section II as $`q_{}=0,\pi `$. This wavevector is measured in units of $`\stackrel{~}{c}^1`$, where $`\stackrel{~}{c}`$ is the rung length, which differs from the reduced lattice constant $`\frac{c}{2}`$. In experiments so far only the basic triplet with $`q_{}=\pi `$ has been observed and the interchange of minima with variation of $`q_z`$ is an effect of the crystallographic lattice geometry and not of the 1D ladder geometry. Excitations with $`q_{}=0`$ are excitations with two dimer quanta and have a minimum energy of twice the gap energy $`\mathrm{\Delta }`$.
In order to obtain results which are quantitatively reliable we have performed series expansions for the 3D coupled system to $`4^{\mathrm{th}}`$ order following the lines described in section II. Because of the complex lattice we did not characterize the clusters which results in a large number of clusters: 5532 clusters in 4th order. The convergence of these series expansion results is excellent for the small values of the expansion parameters $`J_{(lmn)}/J0.4`$ in the application to $`\mathrm{KCuCl}_3`$.
In Figs. 8 and 9 we show results for dispersions along typical lines in $`\stackrel{}{q}`$space together with the data points from ref. . Dispersions are plotted only for the branch which has nonvanishing dipole transition amplitude in lowest order leading to $`+`$ or $``$ sign in Eq. 14 as discussed above. We assume that the effective dimer exchange takes the values determined in previous work and have therefore fixed the following combinations of exchange parameters:
$`J_{(201)}^{}`$ $`=`$ $`2J_{(201)}^{\mathrm{eff}}=0.188,`$ (18)
$`2J_{100}J_{(100)}^{}`$ $`=`$ $`2J_{(100)}^{\mathrm{eff}}=0.110,`$ (19)
$`J_{(1\frac{1}{2}\frac{1}{2})}J_{(1\frac{1}{2}\frac{1}{2})}^{}`$ $`=`$ $`2J_{(1\frac{1}{2}\frac{1}{2})}^{\mathrm{eff}}=0.160`$ (20)
The interactions $`J_{(0,\frac{1}{2},\frac{1}{2})},J_{(0,\frac{1}{2},\frac{1}{2})}^{}`$ are found to be negligibly small. In order to demonstrate the relevance of the individual spin exchange parameters (as opposed to the effective description) we discuss the separations $`(100)`$ and $`(1\pm \frac{1}{2}\frac{1}{2})`$ independently: In Fig. 8 we present the variation of the dispersions $`\omega (\stackrel{}{q})`$ in selected directions in $`\stackrel{}{q}`$ space for different values of the coupling $`J_{(1\frac{1}{2}\frac{1}{2})}=0,0.2,0.4`$ with the remaining parameters fixed as given above. It is seen that the distribution of the effective dimer interaction between parallel and diagonal terms essentially shifts the dispersion curve by constant amounts. A comparison to the corresponding neutron scattering results leads to the conclusion that $`J_{(1\frac{1}{2}\frac{1}{2})}=0.200`$ is the most likely value. The analogous results for different values of the diagonal (zig-zag) coupling $`J_{(100)}^{}=0.1,0.3,0.5`$ are shown in Figs. 9. Here the frequency $`\omega (q_x=0)`$ (which is the energy gap for $`q_z=\pi `$ and the dip energy for $`q_z=0`$) depends only on the effective interaction, whereas the frequency $`\omega (q_x=\pi )`$ (which is minimum for $`q_z=0`$ and a dip energy for $`q_z=2\pi `$) allows to determine the exchange between individual spins.
Comparison of our series expansion results to the neutron scattering data then leads to the following values for the exchange constants between individual spins, not yet determined by the values for the effective dimer interactions published so far:
$`J_{(100)}`$ $``$ $`0.005`$
$`J_{(100)}^{}`$ $``$ $`0.100.`$
These values imply that the ladder system in $`\mathrm{KCuCl}_3`$ is much closer to an alternating spin chain than was believed so far and the leg interaction tends to be ferromagnetic if it is nonzero at all. The results for the interchain interactions in ($`1\frac{1}{2}\frac{1}{2}`$) direction are much less conclusive. The most likely values are
$`J_{(1\frac{1}{2}\frac{1}{2})}`$ $``$ $`0.200`$
$`J_{(1\frac{1}{2}\frac{1}{2})}^{}`$ $``$ $`0.040.`$
However, the error is large and data actually may be compatible also with smaller values for $`J_{(1\frac{1}{2}\frac{1}{2})}`$.
## IV Conclusions
We have investigated the dispersion curve for the low energy triplet excitations of one-dimensional and of weakly coupled zig-zag chains starting from the limit of noninteracting dimers and performing an expansion in the interdimer interactions. The series up to $`10^{\mathrm{th}}`$ order in the 1D case and up to $`4^{\mathrm{th}}`$ order in the 3D case were evaluated explicitly after implementation on a work station. In the 1D case the dispersion curves agree with those obtained from exact diagonalization using the Lanczos algorithm for a large regime around the dimer point; in this regime the method provides a reliable approach to calculate the dispersion in its continuous dependence on the wavevector. In a narrow regime close to the Shastry-Sutherland line we find the minimum of the dispersion curve at an intermediate wavevector $`k_{min},0<k_{min}<\pi `$. As an application of the 3D case we present the application of the method to the $`\mathrm{KCuCl}_3`$structure. Fitting to the dispersion as measured in inelastic neutron scattering experiments we determine the exchange interchange interactions between individual spins in addition to the effective interaction between dimers determined before. It is shown that the effective dimer approximation, when treated in the random phase approximation, sums up the leading powers of the dimer expansion.
## Acknowledgement
We gratefully acknowledge useful discussions with A. Kolezhuk and with U. Neugebauer, who participated in the early stage of this work. We are grateful to N. Cavadini for correspondance and for communicating the data points of ref. in detail. The work was supported by the German Ministry for Research and Technology (BMBF) under contract No. 03Mi5HAN5. |
warning/0003/hep-th0003095.html | ar5iv | text | # The Rest-Frame Instant Form of Relativistic Perfect Fluids with Equation of State ๐=๐โข(๐,๐ ) and of Non-Dissipative Elastic Materials.
## I Introduction
Stability of stellar models for rotating stars, gravity-fluid models, neutron stars, accretion discs around compact objects, collapse of stars, merging of compact objects are only some of the many topics in astrophysics and cosmology in which relativistic hydrodynamics is the basic underlying theory. This theory is also needed in heavy-ions collisions.
As shown in Ref. there are many ways to describe relativistic perfect fluids by means of action functionals both in special and general relativity. Usually, besides the thermodynamical variables $`n`$ (particle number density), $`\rho `$ (energy density), $`p`$ (pressure), $`T`$ (temperature), $`s`$ (entropy per particle), which are spacetime scalar fields whose values represent measurements made in the rest frame of the fluid (Eulerian observers), one characterizes the fluid motion by its unit timelike 4-velocity vector field $`U^\mu `$ (see Appendix A for a review of the relations among the local thermodynamical variables and Appendix B for a review of covariant relativistic thermodynamics following Ref.). However, these variables are constrained due to \[we use a general relativistic notation: โ$`;\mu `$โ denotes a covariant derivative\] i) particle number conservation, $`(nU^\mu )_{;\mu }=0`$; ii) absence of entropy exchange between neighbouring flow lines, $`(nsU^\mu )_{;\mu }=0`$; iii) the requirement that the fluid flow lines should be fixed on the boundary. Therefore one needs Lagrange multipliers to incorporate i) anf ii) into the action and this leads to use Clebsch (or velocity-potential) representations of the 4-velocity and action functionals depending on many redundant variables, generating first and second class constraints at the Hamiltonian level (see Appendix A).
Following Ref. the previous constraint iii) may be enforced by replacing the unit 4-velocity $`U^\mu `$ with a set of spacetime scalar fields $`\stackrel{~}{\alpha }^i(z)`$, $`i=1,2,3,`$ interpreted as โLagrangian (or comoving) coordinates for the fluidโ labelling the fluid flow lines (physically determined by the average particle motions) passing through the points inside the boundary (on the boundary they are fixed: either the $`\stackrel{~}{\alpha }^i(z^o,\stackrel{}{z})`$โs have a compact boundary $`V_\alpha (z^o)`$ or they have assigned boundary conditions at spatial infinity). This requires the choice of an arbitrary spacelike hypersurface on which the $`\alpha ^i`$โs are the 3-coordinates. A similar point of view is contained in the concept of โmaterial spaceโ of Refs., describing the collection of all the idealized points of the material; besides to non-dissipative isentropic fluids the scheme can be applied to isotropic elastic media and anisotropic (crystalline) materials, namely to an arbitrary non-dissipative relativistic continuum . See Ref. for the study of the transformation from Eulerian to Lagrangian coordinates (in the non-relativistic framework of the Euler-Newton equations).
Notice that the use of Lagrangian (comoving) coordinates in place of Eulerian quantities allows the use of standard Poisson brackets in the Hamiltonian description, avoiding the formulation with Lie-Poisson brackets of Ref., which could be recovered by a so-called Lagrangian to Eulerian map.
Let $`M^4`$ be a curved globally hyperbolic spacetime \[with signature $`ฯต(+)`$, $`ฯต=\pm `$\] whose points have locally coordinates $`z^\mu `$. Let $`{}_{}{}^{4}g_{\mu \nu }^{}(z)`$ be its 4-metric with determinant $`{}_{}{}^{4}g=|det{}_{}{}^{4}g_{\mu \nu }^{}|`$. Given a perfect fluid with Lagrangian coordinates $`\stackrel{~}{\alpha }(z)=\{\stackrel{~}{\alpha }^i(z)\}`$, unit 4-velocity vector field $`U^\mu (z)`$ and particle number density $`n(z)`$, let us introduce the number flux vector
$$n(z)U^\mu (z)=\frac{J^\mu (\stackrel{~}{\alpha }^i(z))}{\sqrt{{}_{}{}^{4}g(z)}},$$
(1)
and the densitized fluid number flux vector or material current \[$`ฯต^{0123}=1/\sqrt{{}_{}{}^{4}g}`$; $`_\alpha (\sqrt{{}_{}{}^{4}g}ฯต^{\mu \nu \rho \sigma })=0`$; $`\eta _{123}(\stackrel{~}{\alpha }^i)`$ describes the orientation of the volume in thematerial space\]
$`J^\mu (\stackrel{~}{\alpha }^i(z))`$ $`=`$ $`\sqrt{{}_{}{}^{4}g}ฯต^{\mu \nu \rho \sigma }\eta _{123}(\stackrel{~}{\alpha }^๐ข(z))_\nu \stackrel{~}{\alpha }^1(z)_\rho \stackrel{~}{\alpha }^2(z)_\sigma \stackrel{~}{\alpha }^3(z),`$ (2)
$``$ $`n(z)={\displaystyle \frac{|J(\stackrel{~}{\alpha }^i(z))|}{\sqrt{{}_{}{}^{4}g}}}=\eta _{123}(\stackrel{~}{\alpha }^i(z)){\displaystyle \frac{\sqrt{ฯต{}_{}{}^{4}g_{\mu \nu }^{}(z)J^\mu (\stackrel{~}{\alpha }^i(z))J^\nu (\stackrel{~}{\alpha }^i(z))}}{\sqrt{{}_{}{}^{4}g(z)}}},`$ (3)
$``$ $`_\mu J^\mu (\stackrel{~}{\alpha }^i(z))=\sqrt{{}_{}{}^{4}g}[n(z)U^\mu (z)]_{;\mu }=0,`$ (4)
$``$ $`J^\mu (\stackrel{~}{\alpha }^i(z))_\mu \stackrel{~}{\alpha }^i(z)=[\sqrt{{}_{}{}^{4}g}nU^\mu ](z)_\mu \stackrel{~}{\alpha }^i(z)=0.`$ (5)
This shows that the fluid flow lines, whose tangent vector field is the fluid 4-velocity timelike vector field $`U^\mu `$, are identified by $`\stackrel{~}{\alpha }^๐ข=const.`$ and that the particle number conservation is automatic. Moreover, if the entropy for particle is a function only of the fluid Lagrangian coordinates, $`s=s(\stackrel{~}{\alpha }^๐ข)`$, the assumed form of $`J^\mu `$ also implies automatically the absence of entropy exchange between neighbouring flow lines, $`(nsU^\mu )_{,\mu }=0`$. Since $`U^\mu _\mu s(\stackrel{~}{\alpha }^i)=0`$, the perfect fluid is locally adiabatic; instead for an isentropic fluid we have $`_\mu s=0`$, namely $`s=const.`$.
Even if in general the timelike vector field $`U^\mu (z)`$ is not surface forming (namely has a non-vanishing vorticity, see for instance Ref.), in each point $`z`$ we can consider the spacelike hypersurface orthogonal to the fluid flow line in that point (namely we split the tangent space $`T_zM^4`$ at $`z`$ in the $`U^\mu (z)`$ direction and in the orthogonal complement) and consider $`\frac{1}{3!}[U^\mu ฯต_{\mu \nu \rho \sigma }dz^\nu dz^\rho dz^\sigma ](z)`$ as the infinitesimal 3-volume on it at $`z`$. Then the 3-form
$$\eta [z]=[\eta _{123}(\stackrel{~}{\alpha })d\stackrel{~}{\alpha }^1d\stackrel{~}{\alpha }^2d\stackrel{~}{\alpha }^3](z)=\frac{1}{3!}n(z)[U^\mu ฯต_{\mu \nu \rho \sigma }dz^\nu dz^\rho dz^\sigma ](z),$$
(6)
may be interpreted as the number of particles in this 3-volume. If V is a volume around $`z`$ on the spacelike hypersurface, then $`_V\eta `$ is the number of particle in V and $`_Vs\eta `$ is the total entropy contained in the flow lines included in the volume V. Note that locally $`\eta _{123}`$ can be set to unity by an appropriate choice of coordinates.
In Ref. it is shown that the action functional
$$S[{}_{}{}^{4}g_{\mu \nu }^{},\stackrel{~}{\alpha }]=d^4z\sqrt{{}_{}{}^{4}g(z)}\rho (\frac{|J(\stackrel{~}{\alpha }^i(z))|}{\sqrt{{}_{}{}^{4}g(z)}},s(\stackrel{~}{\alpha }^๐ข(z))),$$
(7)
has a variation with respect to the 4-metric, which gives rise to the correct stress tensor $`T^{\mu \nu }=(\rho +p)U^\mu U^\nu ฯตp{}_{}{}^{4}g_{}^{\mu \nu }`$ with $`p=n\frac{\rho }{n}|_s\rho `$ for a perfect fluid \[see Appendix A\].
The Euler-Lagrange equations associated to the variation of the Lagrangian coordinates are \[$`V_\mu =\mu U_\mu `$ is the Taub current, see Appendix A\]
$`{\displaystyle \frac{1}{\sqrt{{}_{}{}^{4}g}}}{\displaystyle \frac{\delta S}{\delta \stackrel{~}{\alpha }^i}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}ฯต^{\mu \nu \rho \sigma }V_{\mu ;\nu }\eta _{ijk}_\rho \stackrel{~}{\alpha }^j_\sigma \stackrel{~}{\alpha }^knT{\displaystyle \frac{s}{\stackrel{~}{\alpha }^i}}=0,`$ (8)
$`{\displaystyle \frac{1}{\sqrt{{}_{}{}^{4}g}}}{\displaystyle \frac{\delta S}{\delta \stackrel{~}{\alpha }^i}}_\mu \stackrel{~}{\alpha }^i`$ $`=`$ $`{\displaystyle \frac{1}{2}}ฯต^{\alpha \beta \gamma \delta }V_{\alpha ;\beta }U^\nu ฯต_{\nu \mu \gamma \delta }T_\mu s=2V_{[\mu ;\nu ]}U^\nu T_\mu s=0.`$ (10)
As shown in Appendix A, these equations together with the entropy exchange constraint imply the Euler equations implied from the conservation of the stress-energy-momentum tensor.
Therefore, with this description the conservation laws are automatically satisfied and the Euler-Lagrange equations are equivalent to the Euler equations. In Minkowski spacetime the conserved particle number is $`๐ฉ=_{V_\alpha (z^o)}d^3zn(z)U^o(z)=_{V_\alpha (z^o)}d^3zJ^o(\stackrel{~}{\alpha }^i(z))`$, while the conserved entropy per particle is $`_{V_\alpha (z^o)}d^3zs(z)n(z)U^o(z)=_{V_\alpha (z^o)}d^3zs(z)J^o(\stackrel{~}{\alpha }^i)`$. Moreover, the conservation laws $`T^{\mu \nu }{}_{,\nu }{}^{}=0`$ will generate the conserved 4-momentum and angular momentum of the fluid.
However, in Ref. there are only some comments on the Hamiltonian description implied by this particular action.
This description of perfect fluids fits naturally with parametrized Minkowski theories for arbitrary isolated relativistic systems \[see Ref. for a review\] on arbitrary spacelike hypersurfaces, leaves of the foliation of Minkowski spacetime $`M^4`$ associated with one of its 3+1 splittings.
Therefore, the aim of this paper is to find the Wigner covariant rest-frame instant form of the dynamics of a perfect fluid, on the special Wigner hyperplanes orthogonal to the total 4-momentum of the fluid. In this way we will get the description of the global rest frame of the fluid as a whole; instead, the 4-velocity vector field $`U^\mu `$ defines the local rest frame in each point of the fluid by means of the projector $`{}_{}{}^{4}g_{}^{\mu \nu }ฯตU^\mu U^\nu `$. This approach will also produce automatically the coupling of the fluid to ADM metric and tetrad gravity with the extra property of allowing a well defined deparametrization of the theory leading to the rest-frame instant form in Minkowski spacetime with Cartesian coordinates when we put equal to zero the Newton constant G . In this paper we will consider the perfect fluid only in Minkowski spacetime, except for some comments on its coupling to gravity.
The starting point is the foliation of Minkowski spacetime $`M^4`$, which is defined by an embedding $`R\times \mathrm{\Sigma }M^4`$, $`(\tau ,\stackrel{}{\sigma })z^\mu (\tau ,\stackrel{}{\sigma })\mathrm{\Sigma }_\tau `$ and with $`\mathrm{\Sigma }`$ an abstract 3-surface diffeomorphic to $`R^3`$, with $`\mathrm{\Sigma }_\tau `$ its copy embedded in $`M^4`$ labelled by the value $`\tau `$ (the scalar mathematical โtimeโ parameter $`\tau `$ labels the leaves of the foliation, $`\stackrel{}{\sigma }`$ are curvilinear coordinates on $`\mathrm{\Sigma }_\tau `$ and $`\sigma ^{\stackrel{ห}{A}}=(\sigma ^\tau =\tau ,\sigma ^{\stackrel{ห}{r}})`$ are $`\mathrm{\Sigma }_\tau `$-adapted holonomic coordinates for $`M^4`$). See Appendix C for the notations on spacelike hypersurfaces.
In this way one gets a parametrized field theory with a covariant 3+1 splitting of Minkowski spacetime and already in a form suited to the transition to general relativity in its ADM canonical formulation (see also Ref., where a theoretical study of this problem is done in curved spacetimes). The price is that one has to add as new independent configuration variables the embedding coordinates $`z^\mu (\tau ,\stackrel{}{\sigma })`$ of the points of the spacelike hypersurface $`\mathrm{\Sigma }_\tau `$ \[the only ones carrying Lorentz indices\] and then to define the fields on $`\mathrm{\Sigma }_\tau `$ so that they know the hypersurface $`\mathrm{\Sigma }_\tau `$ of $`\tau `$-simultaneity \[for a Klein-Gordon field $`\varphi (x)`$, this new field is $`\stackrel{~}{\varphi }(\tau ,\stackrel{}{\sigma })=\varphi (z(\tau ,\stackrel{}{\sigma }))`$: it contains the non-local information about the embedding\]. Then one rewrites the Lagrangian of the given isolated system in the form required by the coupling to an external gravitational field, makes the previous 3+1 splitting of Minkowski spacetime and interpretes all the fields of the system as the new fields on $`\mathrm{\Sigma }_\tau `$ (they are Lorentz scalars, having only surface indices). Instead of considering the 4-metric as describing a gravitational field (and therefore as an independent field as it is done in metric gravity, where one adds the Hilbert action to the action for the matter fields), here one replaces the 4-metric with the the induced metric $`g_{\stackrel{ห}{A}\stackrel{ห}{B}}[z]=z_{\stackrel{ห}{A}}^\mu \eta _{\mu \nu }z_{\stackrel{ห}{B}}^\nu `$ on $`\mathrm{\Sigma }_\tau `$ \[a functional of $`z^\mu `$; $`z_{\stackrel{ห}{A}}^\mu =z^\mu /\sigma ^{\stackrel{ห}{A}}`$ are flat inverse tetrad fields on Minkowski spacetime with the $`z_{\stackrel{ห}{r}}^\mu `$โs tangent to $`\mathrm{\Sigma }_\tau `$\] and considers the embedding coordinates $`z^\mu (\tau ,\stackrel{}{\sigma })`$ as independent fields \[this is not possible in metric gravity, because in curved spacetimes $`z_{\stackrel{ห}{A}}^\mu z^\mu /\sigma ^{\stackrel{ห}{A}}`$ are not tetrad fields so that holonomic coordinates $`z^\mu (\tau ,\stackrel{}{\sigma })`$ do not exist\]. From this Lagrangian, besides a Lorentz-scalar form of the constraints of the given system, we get four extra primary first class constraints
$$_\mu (\tau ,\stackrel{}{\sigma })=\rho _\mu (\tau ,\stackrel{}{\sigma })l_\mu (\tau ,\stackrel{}{\sigma })T_{sys}^{\tau \tau }(\tau ,\stackrel{}{\sigma })z_{\stackrel{ห}{r}\mu }(\tau ,\stackrel{}{\sigma })T_{sys}^{\stackrel{ห}{r}\tau }(\tau ,\stackrel{}{\sigma })0,$$
(11)
\[ here $`T_{sys}^{\tau \tau }(\tau ,\stackrel{}{\sigma })=(\tau ,\stackrel{}{\sigma })`$, $`T_{sys}^{\stackrel{ห}{r}\tau }(\tau ,\stackrel{}{\sigma })=^{\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })`$, are the components of the energy-momentum tensor in the holonomic coordinate system, corresponding to the energy- and momentum-density of the isolated system; one has $`\{_{(\mu )}(\tau ,\stackrel{}{\sigma }),_{(\nu )}(\tau ,\stackrel{}{\sigma }^{^{}})\}=0`$\] implying the independence of the description from the choice of the 3+1 splitting, i.e. from the choice of the foliation with spacelike hypersufaces. As shown in Appendix C the evolution vector is given by $`z_\tau ^\mu =N_{[z](flat)}l^\mu +N_{[z](flat)}^{\stackrel{ห}{r}}z_{\stackrel{ห}{r}}^\mu `$, where $`l^\mu (\tau ,\stackrel{}{\sigma })`$ is the normal to $`\mathrm{\Sigma }_\tau `$ in $`z^\mu (\tau ,\stackrel{}{\sigma })`$ and $`N_{[z](flat)}(\tau ,\stackrel{}{\sigma })`$, $`N_{[z](flat)}^{\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })`$ are the flat lapse and shift functions defined through the metric like in general relativity: however, now they are not independent variables but functionals of $`z^\mu (\tau ,\stackrel{}{\sigma })`$.
The Dirac Hamiltonian contains the piece $`d^3\sigma \lambda ^\mu (\tau ,\stackrel{}{\sigma })_\mu (\tau ,\stackrel{}{\sigma })`$ with $`\lambda ^\mu (\tau ,\stackrel{}{\sigma })`$ Dirac multipliers. It is possible to rewrite the integrand in the form \[$`\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}=ฯต{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}`$ is the inverse of the spatial metric $`g_{\stackrel{ห}{r}\stackrel{ห}{s}}={}_{}{}^{4}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}=ฯต{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$, with $`{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ of positive signature $`(+++)`$\]
$`\lambda _\mu (\tau ,\stackrel{}{\sigma })^\mu (\tau ,\stackrel{}{\sigma })`$ $`=`$ $`[(\lambda _\mu l^\mu )(l_\nu ^\nu )(\lambda _\mu z_r^\mu )(\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}z_{\stackrel{ห}{s}\nu }^\nu )](\tau ,\stackrel{}{\sigma })\stackrel{def}{=}`$ (12)
$`\stackrel{def}{=}`$ $`N_{(flat)}(\tau ,\stackrel{}{\sigma })(l_\mu ^\mu )(\tau ,\stackrel{}{\sigma })N_{(flat)\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })(\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}z_{\stackrel{ห}{s}\nu }^\nu )(\tau ,\stackrel{}{\sigma }),`$ (13)
with the (non-holonomic form of the) constraints $`(l_\mu ^\mu )(\tau ,\stackrel{}{\sigma })0`$, $`(\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}z_{\stackrel{ห}{s}\mu }^\mu )(\tau ,\stackrel{}{\sigma })0`$, satisfying the universal Dirac algebra of the ADM constraints. In this way we have defined new flat lapse and shift functions
$`N_{(flat)}(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`\lambda _\mu (\tau ,\stackrel{}{\sigma })l^\mu (\tau ,\stackrel{}{\sigma }),`$ (14)
$`N_{(flat)\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`\lambda _\mu (\tau ,\stackrel{}{\sigma })z_{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma }).`$ (15)
which have the same content of the arbitrary Dirac multipliers $`\lambda _\mu (\tau ,\stackrel{}{\sigma })`$, namely they multiply primary first class constraints satisfying the Dirac algebra. In Minkowski spacetime they are quite distinct from the previous lapse and shift functions $`N_{[z](flat)}`$, $`N_{[z](flat)\stackrel{ห}{r}}`$, defined starting from the metric. Since the Hamilton equations imply $`z_\tau ^\mu (\tau ,\stackrel{}{\sigma })=\lambda ^\mu (\tau ,\stackrel{}{\sigma })`$, it is only through the equations of motion that the two types of functions are identified. Instead in general relativity the lapse and shift functions defined starting from the 4-metric are the coefficients (in the canonical part $`H_c`$ of the Hamiltonian) of secondary first class constraints satisfying the Dirac algebra independently from the equations of motion.
For the relativistic perfect fluid with equation of state $`\rho =\rho (n,s)`$ in Minkowski spacetime, we have only to replace the external 4-metric $`{}_{}{}^{4}g_{\mu \nu }^{}`$ with $`g_{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })={}_{}{}^{4}g_{\stackrel{ห}{A}\stackrel{ห}{B}}^{}(\tau ,\stackrel{}{\sigma })`$ and the scalar fields for the Lagrangian coordinates with $`\alpha ^i(\tau ,\stackrel{}{\sigma })=\stackrel{~}{\alpha }^i(z(\tau ,\stackrel{}{\sigma }))`$; now either the $`\alpha ^i(\tau ,\stackrel{}{\sigma })`$โs have a compact boundary $`V_\alpha (\tau )\mathrm{\Sigma }_\tau `$ or have boundary conditions at spatial infinity. For each value of $`\tau `$, one could invert $`\alpha ^i=\alpha ^i(\tau ,\stackrel{}{\sigma })`$ to $`\stackrel{}{\sigma }=\stackrel{}{\sigma }(\tau ,\alpha ^i)`$ and use the $`\alpha ^i`$โs as a special coordinate system on $`\mathrm{\Sigma }_\tau `$ inside the support $`V_\alpha (\tau )\mathrm{\Sigma }_\tau `$: $`z^\mu (\tau ,\stackrel{}{\sigma }(\tau ,\alpha ^i))=\stackrel{ห}{z}^\mu (\tau ,\alpha ^i)`$.
By going to $`\mathrm{\Sigma }_\tau `$-adapted coordinates such that $`\eta _{123}(\alpha )=1`$ we get \[$`\gamma =|detg_{\stackrel{ห}{r}\stackrel{ห}{s}}|`$; $`\sqrt{g}=\sqrt{{}_{}{}^{4}g}=\sqrt{|detg_{\stackrel{ห}{A}\stackrel{ห}{B}}|}=N\sqrt{\gamma }`$\]
$`J^{\stackrel{ห}{A}}(\alpha ^i(\tau ,\stackrel{}{\sigma }))`$ $`=`$ $`[N\sqrt{\gamma }nU^{\stackrel{ห}{A}}](\tau ,\stackrel{}{\sigma }),`$ (16)
$`J^\tau (\alpha ^i(\tau ,\stackrel{}{\sigma }))`$ $`=`$ $`[ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{r}}\alpha ^1_{\stackrel{ห}{u}}\alpha ^2_{\stackrel{ห}{v}}\alpha ^3](\tau ,\stackrel{}{\sigma })=det(_{\stackrel{ห}{r}}\alpha ^i)(\tau ,\stackrel{}{\sigma }),`$ (18)
$`J^{\stackrel{ห}{r}}(\alpha ^i(\tau ,\stackrel{}{\sigma }))`$ $`=`$ $`[{\displaystyle \underset{i=1;i,j,kcyclic}{\overset{3}{}}}_\tau \alpha ^iฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k](\tau ,\stackrel{}{\sigma })=`$ (20)
$`={\displaystyle \frac{1}{2}}ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}ฯต_{ijk}[_\tau \alpha ^i_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k](\tau ,\stackrel{}{\sigma }),`$
$``$ $`n(\tau ,\stackrel{}{\sigma })={\displaystyle \frac{|J|}{N\sqrt{\gamma }}}(\tau ,\stackrel{}{\sigma })={\displaystyle \frac{\sqrt{ฯตg_{\stackrel{ห}{A}\stackrel{ห}{B}}J^{\stackrel{ห}{A}}J^{\stackrel{ห}{B}}}}{N\sqrt{\gamma }}}(\tau ,\stackrel{}{\sigma }),`$ (22)
with $`๐ฉ=_{V_\alpha (\tau )}d^3\sigma J^\tau (\alpha ^i(\tau ,\stackrel{}{\sigma }))`$ giving the conserved particle number and $`_{V_\alpha (\tau )}d^3\sigma (sJ^\tau )(\tau ,\stackrel{}{\sigma })`$ giving the conserved entropy per particle.
The action becomes
$`S`$ $`=`$ $`{\displaystyle ๐\tau d^3\sigma L(z^\mu (\tau ,\stackrel{}{\sigma }),\alpha ^i(\tau ,\stackrel{}{\sigma }))}=`$ (23)
$`=`$ $`{\displaystyle ๐\tau d^3\sigma (N\sqrt{\gamma })(\tau ,\stackrel{}{\sigma })\rho (\frac{|J(\alpha ^i(\tau ,\stackrel{}{\sigma }))|}{(N\sqrt{\gamma })(\tau \stackrel{}{\sigma })},s(\alpha ^i(\tau ,\stackrel{}{\sigma })))}=`$ (24)
$`=`$ $`{\displaystyle ๐\tau d^3\sigma (N\sqrt{\gamma })(\tau ,\stackrel{}{\sigma })}`$ (26)
$`\rho ({\displaystyle \frac{1}{\sqrt{\gamma (\tau ,\stackrel{}{\sigma })}}}\sqrt{\left[(J^\tau )^2{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}{\displaystyle \frac{J^{\stackrel{ห}{u}}+N^{\stackrel{ห}{u}}J^\tau }{N}}{\displaystyle \frac{J^{\stackrel{ห}{v}}+N^{\stackrel{ห}{v}}J^\tau }{N}}\right](\tau ,\stackrel{}{\sigma };\alpha ^i(\tau ,\stackrel{}{\sigma }))},s(\alpha ^i(\tau ,\stackrel{}{\sigma }))),`$
with $`N=N_{[z](flat)}`$, $`N^{\stackrel{ห}{r}}=N_{[z](flat)}^{\stackrel{ห}{r}}`$.
This is the form of the action whose Hamiltonian formulation will be studied in this paper.
We shall begin in Section II with the simple case of dust, whose equation of state is $`\rho =\mu n`$.
In Section III we will define the โexternalโ and โinternalโ centers of mass of the dust.
In Section IV we will study Dixonโs multipoles of a perfect fluid on the Wigner hyperplane in Minkowski spacetime using the dust as an example.
Then in Section V we will consider some equations of state for isentropic fluids and we will make some comments on non-isentropic fluids.
In Section VI we will define the coupling to ADM metric and tetrad gravity.
In Section VII we will describe with the same technology isentropic elastic media.
In the Conclusions, after some general remarks, we will delineate the treatment of perfect fluids in tetrad gravity (this will be the subject of a future paper).
In Appendix A there is a review of some of the results of Ref. for relativistic perfect fluids.
In Appendix B there is a review of covariant relativistic thermodynamics of equilibrium and non-equilibrium.
In Appendix C there is some notation on spacelike hypersurfaces.
In Appendix D there is the definition of other types of Dixonโs multipoles.
## II Dust.
Let us consider first the simplest case of an isentropic perfect fluid, a dust with $`p=0`$, $`s=const.`$, and equation of state $`\rho =\mu n`$. In this case the chemical potential $`\mu `$ is the rest mass-energy of a fluid particle: $`\mu =m`$ (see Appendix A).
Eq.(LABEL:I10) implies that the Lagrangian density is \[we shall use the notation $`g_{\stackrel{ห}{A}\stackrel{ห}{B}}={}_{}{}^{4}g_{\stackrel{ห}{A}\stackrel{ห}{B}}^{}`$ with signature $`ฯต(+)`$, $`ฯต=\pm 1`$; by using the notation with lapse and shift functions given in Appendix C we get: $`g_{\tau \tau }=ฯต(N^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}})`$, $`g_{\tau \stackrel{ห}{r}}=ฯต{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}N^{\stackrel{ห}{s}}`$, $`g_{\stackrel{ห}{r}\stackrel{ห}{s}}=ฯต{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ with $`{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ of positive signature (+++), $`g^{\tau \tau }=\frac{ฯต}{N^2}`$, $`g^{\tau \stackrel{ห}{r}}=ฯต\frac{N^{\stackrel{ห}{r}}}{N^2}`$, $`g^{\stackrel{ห}{r}\stackrel{ห}{s}}=ฯต({}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}\frac{N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}}}{N^2})`$; the inverse of the spatial 4-metric $`{}_{}{}^{4}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ is denoted $`\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}={}_{}{}^{4}\gamma _{}^{\stackrel{ห}{r}\stackrel{ห}{s}}=ฯต{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}`$, where $`{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}`$ is the inverse of the 3-metric $`{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ and we use $`\sqrt{\gamma }=\sqrt{{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}}`$\]
$`L(\alpha ^i,z^\mu )`$ $`=`$ $`\sqrt{g}\rho =\mu \sqrt{g}n=\mu \sqrt{ฯตg_{\stackrel{ห}{A}\stackrel{ห}{B}}J^{\stackrel{ห}{A}}J^{\stackrel{ห}{B}}}=`$ (28)
$`=`$ $`\mu N\sqrt{(J^\tau )^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{r}}Y^{\stackrel{ห}{s}}}=\mu NX,`$ (29)
$`Y^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \frac{1}{N}}(J^{\stackrel{ห}{r}}+N^{\stackrel{ห}{r}}J^\tau ),`$ (31)
$`X`$ $`=`$ $`\sqrt{(J^\tau )^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{r}}Y^{\stackrel{ห}{s}}}={\displaystyle \frac{\sqrt{g}}{N}}n=\sqrt{\gamma }n,`$ (32)
with $`J^\tau `$, $`J^{\stackrel{ห}{r}}`$ given in Eqs.(22).
The momentum conjugate to $`\alpha ^i`$ is
$`\mathrm{\Pi }_i`$ $`=`$ $`{\displaystyle \frac{L}{_\tau \alpha ^i}}=\mu {\displaystyle \frac{Y^{\stackrel{ห}{t}}{}_{}{}^{3}g_{\stackrel{ห}{t}\stackrel{ห}{r}}^{}ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k}{X}}|_{i,j,kcyclic}=`$ (33)
$`=`$ $`\mu {\displaystyle \frac{Y^{\stackrel{ห}{t}}}{2X}}{}_{}{}^{3}g_{\stackrel{ห}{t}\stackrel{ห}{r}}^{}ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}ฯต_{ijk}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k=\mu {\displaystyle \frac{Y^{\stackrel{ห}{r}}}{X}}T_{\stackrel{ห}{r}i},`$ (34)
, (35)
$`T_{\stackrel{ห}{t}i}`$ $`\stackrel{def}{=}`$ $`{\displaystyle \frac{1}{2}}g_{\stackrel{ห}{t}\stackrel{ห}{r}}ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}ฯต_{ijk}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k=g_{\stackrel{ห}{t}\stackrel{ห}{r}}(adJ_{i\stackrel{ห}{r}}),`$ (36)
where $`adJ_{i\stackrel{ห}{r}}=(detJ)J_{i\stackrel{ห}{r}}^1`$ is the adjoint matrix of the Jacobian $`J=(J_{i\stackrel{ห}{r}}=_{\stackrel{ห}{r}}\alpha ^i)`$ of the transformation from the Lagrangian coordinates $`\alpha ^i(\tau ,\stackrel{}{\sigma })`$ to the Eulerian ones $`\stackrel{}{\sigma }`$ on $`\mathrm{\Sigma }_\tau `$.
The momentum conjugate to $`z^\mu `$ is
$$\rho _\mu (\tau ,\stackrel{}{\sigma })=\frac{L}{z_\tau ^\mu }(\tau ,\stackrel{}{\sigma })=\left[\mu l_\mu \frac{(J^\tau )^2}{X}+\mu z_{r\mu }J^\tau \frac{Y^r}{X}\right](\tau ,\stackrel{}{\sigma }).$$
(37)
The following Poisson brackets are assumed
$`\{z^\mu (\tau ,\stackrel{}{\sigma }),\rho _\nu (\tau ,\stackrel{}{\sigma }^{^{}}\}=\eta _\nu ^\mu \delta ^3(\stackrel{}{\sigma }\stackrel{}{\sigma }^{^{}}),`$ (38)
$`\{\alpha ^i(\tau ,\stackrel{}{\sigma }),\mathrm{\Pi }_j(\tau ,\stackrel{}{\sigma }^{^{}})\}=\delta _j^i\delta ^3(\stackrel{}{\sigma }\stackrel{}{\sigma }^{^{}}).`$ (39)
We can express $`Y^{\stackrel{ห}{r}}/X`$ in terms of $`\mathrm{\Pi }_i`$ with the help of the inverse $`(T^1)^{\stackrel{ห}{r}i}`$ of the matrix $`T_{\stackrel{ห}{t}i}`$
$$\frac{Y^{\stackrel{ห}{r}}}{X}=\frac{1}{\mu }(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i,$$
(40)
where
$$(T^1)^{\stackrel{ห}{r}i}=\frac{{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}_{\stackrel{ห}{s}}\alpha ^i}{det(_{\stackrel{ห}{u}}\alpha ^k)}.$$
(41)
From the definition of $`X`$ we find
$$X=\frac{\mu J^\tau }{\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_i\mathrm{\Pi }_j}}.$$
(42)
Consequently, we can get the expression of the velocities of the Lagrangian coordinates in terms of the momenta
$$_\tau \alpha ^i=\frac{J^{\stackrel{ห}{r}}_{\stackrel{ห}{r}}\alpha ^i}{J^\tau }=\frac{(N^{\stackrel{ห}{r}}J^\tau NY^{\stackrel{ห}{r}})_{\stackrel{ห}{r}}\alpha ^i}{J^\tau },$$
(43)
namely
$$_\tau \alpha ^i=_{\stackrel{ห}{r}}\alpha ^i\left[N^{\stackrel{ห}{r}}N(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right].$$
(44)
Now $`\rho _\mu `$ can be expressed as a function of the $`z`$โs, $`\alpha `$โs and $`\mathrm{\Pi }`$โs:
$$\rho _\mu =l_\mu J^\tau \sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j}+z_{\stackrel{ห}{r}\mu }J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i.$$
(45)
Since the Lagrangian is homogenous in the velocities, the Hamiltonian is only
$$H_D=d^3\sigma \lambda ^\mu (\tau ,\stackrel{}{\sigma })_\mu (\tau ,\stackrel{}{\sigma }),$$
(46)
where the $`_\mu `$ are the primary constraints
$`_\mu `$ $`=`$ $`\rho _\mu l_\mu +z_{\stackrel{ห}{r}\mu }^{\stackrel{ห}{R}}0,`$ (47)
$``$ $`=`$ $`T^{\tau \tau }=J^\tau \sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j},`$ (49)
$`^{\stackrel{ห}{r}}`$ $`=`$ $`T^{\tau \stackrel{ห}{r}}=J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i.`$ (50)
satisfying
$$\{_\mu (\tau ,\stackrel{}{\sigma }),_\nu (\tau ,\stackrel{}{\sigma }^{^{}})\}=0.$$
(51)
One finds that $`\{_\mu (\tau ,\stackrel{}{\sigma }),H_D\}=0`$. Therefore, there are only the four first class constraints $`_\mu (\tau ,\stackrel{}{\sigma })0`$. They describe the arbitrariness of the foliation: physical results do not depend on its choice.
The conserved Poincarรฉ generators are (the suffix โsโ denotes the hypersurface $`\mathrm{\Sigma }_\tau `$)
$`p_s^\mu ={\displaystyle d^3\sigma \rho ^\mu (\tau ,\stackrel{}{\sigma })},`$ (52)
$`J_s^{\mu \nu }={\displaystyle d^3\sigma [z^\mu (\tau ,\stackrel{}{\sigma })\rho ^\nu (\tau ,\stackrel{}{\sigma })z^\nu (\tau ,\stackrel{}{\sigma })\rho ^\mu (\tau ,\stackrel{}{\sigma })]},`$ (53)
and one has
$$\{z^\mu (\tau ,\stackrel{}{\sigma }),p_s^\nu \}=\eta ^{\mu \nu },$$
(54)
$`{\displaystyle d^3\sigma _\mu (\tau ,\stackrel{}{\sigma })}`$ $`=`$ $`p_s^\mu {\displaystyle d^3\sigma \left[l_\mu J^\tau \sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma })}+`$ (55)
$`+`$ $`{\displaystyle d^3\sigma \left[z_{\stackrel{ห}{r}\mu }J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i\right](\tau ,\stackrel{}{\sigma })}0.`$ (56)
Let us now restrict ourselves to spacelike hyperplanes $`\mathrm{\Sigma }_\tau `$ by imposing the gauge-fixings
$`\zeta ^\mu (\tau ,\stackrel{}{\sigma })`$ $`=`$ $`z^\mu (\tau ,\stackrel{}{\sigma })x_s^\mu (\tau )b_{\stackrel{ห}{r}}^\mu (\tau )\sigma ^{\stackrel{ห}{r}}0,`$ (59)
$`\{\zeta ^\mu (\tau ,\stackrel{}{\sigma }),_\nu (\tau ,\stackrel{}{\sigma }^{^{}})\}=\eta _\nu ^\mu \delta ^3(\stackrel{}{\sigma }\stackrel{}{\sigma }^{^{}}),`$
where $`x_s^\mu (\tau )`$ is an arbitrary point of $`\mathrm{\Sigma }_\tau `$, chosen as origin of the coordinates $`\sigma ^{\stackrel{ห}{r}}`$, and $`b_{\stackrel{ห}{r}}^\mu (\tau )`$, $`\stackrel{ห}{r}=1,2,3`$, are three orthonormal vectors such that the constant (future pointing) normal to the hyperplane is
$$l^\mu (\tau ,\stackrel{}{\sigma })l^\mu =b_\tau ^\mu =ฯต^\mu {}_{\alpha \beta \gamma }{}^{}b_{\stackrel{ห}{1}}^{\alpha }(\tau )b_{\stackrel{ห}{2}}^\beta (\tau )b_{\stackrel{ห}{3}}^\gamma (\tau ).$$
(60)
Therefore, we get
$`z_{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma })b_{\stackrel{ห}{r}}^\mu (\tau ),`$ (61)
$`z_\tau ^\mu (\tau ,\stackrel{}{\sigma })\dot{x}_s^\mu (\tau )+\dot{b}_{\stackrel{ห}{r}}^\mu (\tau )\sigma ^{\stackrel{ห}{r}},`$ (62)
$`g_{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })ฯต\delta _{\stackrel{ห}{r}\stackrel{ห}{s}},\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })ฯต\delta ^{\stackrel{ห}{r}\stackrel{ห}{s}},\gamma (\tau ,\stackrel{}{\sigma })1.`$ (63)
By introducing the Dirac brackets for the resulting second class constraints
$$\{A,B\}{}_{}{}^{}=\{A,B\}d^3\sigma [\{A,\zeta ^\mu (\tau ,\stackrel{}{\sigma })\}\{_\mu (\tau ,\stackrel{}{\sigma }),B\}\{A,_\mu (\tau ,\stackrel{}{\sigma })\}\{\zeta ^\mu (\tau ,\stackrel{}{\sigma }),B\}],$$
(64)
we find that, by using Eq.(54) and (56) \[with $`x_s^\mu (\tau )=z^\mu (\tau ,\stackrel{}{\sigma })b_{\stackrel{ห}{r}}^\mu (\tau )\sigma ^{\stackrel{ห}{r}}\zeta ^\mu (\tau ,\stackrel{}{\sigma })`$ and with the assumption $`\{b_{\stackrel{ห}{r}}^\mu (\tau ),p_s^\nu \}=0`$\], we get
$$\{x_s^\mu (\tau ),p_s^\nu (\tau )\}{}_{}{}^{}=\eta ^{\mu \nu }.$$
(65)
The ten degrees of freedom describing the hyperplane are $`x_s^\mu (\tau )`$ with conjugate momentum $`p_s^\mu `$ and six variables $`\varphi _\lambda (\tau )`$, $`\lambda =1,..,6`$, which parametrize the orthonormal tetrad $`b_{\stackrel{ห}{A}}^\mu (\tau )`$, with their conjugate momenta $`T_\lambda (\tau )`$.
The preservation of the gauge-fixings $`\zeta ^\mu (\tau ,\stackrel{}{\sigma })0`$ in time implies
$$\frac{d}{d\tau }\zeta ^\mu (\tau ,\stackrel{}{\sigma })=\{\zeta ^\mu (\tau ,\stackrel{}{\sigma }),H_D\}=\lambda ^\mu (\tau ,\stackrel{}{\sigma })\dot{x}_s^\mu (\tau )\dot{b}_{\stackrel{ห}{r}}^\mu (\tau )\sigma ^{\stackrel{ห}{r}}0,$$
(66)
so that one has \[by using $`\dot{b}_\tau ^\mu =0`$ and $`\dot{b}_{\stackrel{ห}{r}}^\mu (\tau )b_{\stackrel{ห}{r}}^\nu (\tau )=b_{\stackrel{ห}{r}}^\mu (\tau )\dot{b}_{\stackrel{ห}{r}}^\nu (\tau )`$\]
$`\lambda ^\mu (\tau ,\stackrel{}{\sigma })`$ $``$ $`\stackrel{~}{\lambda }^\mu (\tau )+\stackrel{~}{\lambda }^\mu {}_{\nu }{}^{}(\tau )b_{\stackrel{ห}{r}}^\nu (\tau )\sigma ^{\stackrel{ห}{r}},`$ (69)
$`\stackrel{~}{\lambda }^\mu (\tau )=\dot{x}_s^\mu (\tau ),`$
$`\stackrel{~}{\lambda }^{\mu \nu }(\tau )=\stackrel{~}{\lambda }^{\nu \mu }(\tau )={\displaystyle \frac{1}{2}}[\dot{b}_{\stackrel{ห}{r}}^\mu (\tau )b_{\stackrel{ห}{r}}^\nu (\tau )b_{\stackrel{ห}{r}}^\mu (\tau )\dot{b}_{\stackrel{ห}{r}}^\nu (\tau )].`$
Thus, the Dirac Hamiltonian becomes
$$H_D=\stackrel{~}{\lambda }^\mu (\tau )\stackrel{~}{}_\mu (\tau )\frac{1}{2}\stackrel{~}{\lambda }^{\mu \nu }(\tau )\stackrel{~}{}_{\mu \nu }(\tau ),$$
(70)
and this shows that the gauge fixings $`\zeta ^\mu (\tau ,\stackrel{}{\sigma })0`$ do not transform completely the constraints $`_\mu (\tau ,\stackrel{}{\sigma })0`$ in their second class partners; still the following ten first class constraints are left
$`\stackrel{~}{}^\mu (\tau )`$ $`=`$ $`{\displaystyle d^3\sigma ^\mu (\tau ,\stackrel{}{\sigma })}=p_s^\mu `$ (71)
$``$ $`{\displaystyle d^3\sigma \left(J^\tau \left[l^\mu \sqrt{\mu ^2+\delta _{\stackrel{ห}{u}\stackrel{ห}{v}}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j}+b_{\stackrel{ห}{r}}^\mu (T^1)^{\stackrel{ห}{r}l}\mathrm{\Pi }_l\right]\right)(\tau ,\stackrel{}{\sigma })}0,`$ (72)
$`\stackrel{~}{}^{\mu \nu }(\tau )`$ $`=`$ $`b_{\stackrel{ห}{r}}^\mu (\tau ){\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}^\nu (\tau ,\stackrel{}{\sigma })}b_{\stackrel{ห}{r}}^\nu (\tau ){\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma })}=`$ (74)
$`=`$ $`S_s^{\mu \nu }(\tau )`$ (75)
$``$ $`[b_{\stackrel{ห}{r}}^\mu (\tau )b_\tau ^\nu b_{\stackrel{ห}{r}}^\nu (\tau )b_\tau ^\mu ]{\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\left(J^\tau \sqrt{\mu ^2+\delta _{\stackrel{ห}{u}\stackrel{ห}{v}}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j}\right)(\tau ,\stackrel{}{\sigma })}+`$ (76)
$`+`$ $`[b_{\stackrel{ห}{r}}^\mu (\tau )b_{\stackrel{ห}{s}}^\nu (\tau )b_{\stackrel{ห}{r}}^\nu (\tau )b_{\stackrel{ห}{s}}^\mu (\tau )]{\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\left(J^\tau (T^1)^{\stackrel{ห}{s}l}\mathrm{\Pi }_l\right)(\tau ,\stackrel{}{\sigma })}0.`$ (77)
Here $`S_s^{\mu \nu }`$ is the spin part of the Lorentz generators
$`J_s^{\mu \nu }`$ $`=`$ $`x_s^\mu p_s^\nu x_s^\nu p_s^\mu +S_s^{\mu \nu },`$ (79)
$`S_s^{\mu \nu }=b_{\stackrel{ห}{r}}^\mu (\tau ){\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\rho ^\nu (\tau ,\stackrel{}{\sigma })}b_{\stackrel{ห}{r}}^\nu (\tau ){\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\rho ^\mu (\tau ,\stackrel{}{\sigma })}.`$
As shown in Ref. instead of finding $`\varphi _\lambda (\tau ),T_\lambda (\tau )`$, one can use the redundant variables $`b_{\stackrel{ห}{A}}^\mu (\tau ),S_s^{\mu \nu }(\tau )`$, with the following Dirac brackets assuring the validity of the orthonormality condition $`\eta ^{\mu \nu }b_{\stackrel{ห}{A}}^\mu \eta ^{\stackrel{ห}{A}\stackrel{ห}{b}}b_{\stackrel{ห}{B}}^\nu =0`$ \[$`C_{\gamma \delta }^{\mu \nu \alpha \beta }=\eta _\gamma ^\nu \eta _\delta ^\alpha \eta ^{\mu \beta }+\eta _\gamma ^\mu \eta _\delta ^\beta \eta ^{\nu \alpha }\eta _\gamma ^\nu \eta _\delta ^\beta \eta ^{\mu \alpha }\eta _\gamma ^\mu \eta _\delta ^\alpha \eta ^{\nu \beta }`$ are the structure constants of the Lorentz group\]
$`\{S_s^{\mu \nu },b_{\stackrel{ห}{A}}^\rho \}{}_{}{}^{}=\eta ^{\rho \nu }b_{\stackrel{ห}{A}}^\mu \eta ^{\rho \mu }b_{\stackrel{ห}{A}}^\nu `$ (80)
$`\{S_s^{\mu \nu },S_s^{\alpha \beta }\}{}_{}{}^{}=C_{\gamma \delta }^{\mu \nu \alpha \beta }S_s^{\gamma \delta },`$ (81)
so that, while $`\stackrel{~}{}^\mu (\tau )0`$ has zero Dirac bracket with itself and with $`\stackrel{~}{}^{\mu \nu }(\tau )0`$, these last six constraints have the Dirac brackets
$$\{\stackrel{~}{}^{\mu \nu }(\tau ),\stackrel{~}{}^{\alpha \beta }(\tau )\}{}_{}{}^{}=C_{\gamma \delta }^{\mu \nu \alpha \beta }\stackrel{~}{}^{\gamma \delta }(\tau )0.$$
(82)
We have now only the variables: $`x_s^\mu `$, $`p_s^\mu `$, $`b_{\stackrel{ห}{A}}^\mu `$, $`S_s^{\mu \nu }`$, $`\alpha ^i`$, $`\mathrm{\Pi }_i`$ with the following Dirac brackets:
$`\{x_s^\mu (\tau ),p_s^\nu (\tau )\}^{}`$ $`=`$ $`\eta ^{\mu \nu },`$ (83)
$`\{S_s^{\mu \nu }(\tau ),b_{\stackrel{ห}{A}}^\rho (\tau )\}^{}`$ $`=`$ $`\eta ^{\rho \nu }b_{\stackrel{ห}{A}}^\mu (\tau )\eta ^{\rho \mu }b_{\stackrel{ห}{A}}^\nu (\tau ),`$ (84)
$`\{S_s^{\mu \nu }(\tau ),S_s^{\alpha \beta }(\tau )\}^{}`$ $`=`$ $`C_{\gamma \delta }^{\mu \nu \alpha \beta }S_s^{\gamma \delta }(\tau ),`$ (85)
$`\{\alpha ^i(\tau ,\stackrel{}{\sigma }),\mathrm{\Pi }_j(\tau ,\stackrel{}{\sigma }^{^{}})\}^{}`$ $`=`$ $`\delta _j^i\delta ^3(\stackrel{}{\sigma }\stackrel{}{\sigma }^{^{}}).`$ (86)
After the restriction to spacelike hyperplanes we have $`z_{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma })b_{\stackrel{ห}{r}}^\mu (\tau )`$, so that $`z_\tau ^\mu (\tau ,\stackrel{}{\sigma })N_{[z](flat)}(\tau ,\stackrel{}{\sigma })l^\mu (\tau ,\stackrel{}{\sigma })+N_{[z](flat)}^{\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })`$ $`b_{\stackrel{ห}{r}}^{(\mu )}(\tau ,\stackrel{}{\sigma })\dot{x}_s^\mu (\tau )+\dot{b}_{\stackrel{ห}{r}}^\mu (\tau )\sigma ^{\stackrel{ห}{r}}=\stackrel{~}{\lambda }^\mu (\tau )\stackrel{~}{\lambda }^{\mu \nu }(\tau )b_{\stackrel{ห}{r}\nu }(\tau )\sigma ^{\stackrel{ห}{r}}`$. As said in the Introduction only now we get the coincidence of the two definitions of flat lapse and shift functions (this point was missed in the older treatments of parametrized Minkowski theories):
$`N_{[z](flat)}(\tau ,\stackrel{}{\sigma })`$ $``$ $`N_{(flat)}(\tau ,\stackrel{}{\sigma })=\stackrel{~}{\lambda }_\mu (\tau )l^\mu l^\mu \stackrel{~}{\lambda }_{\mu \nu }(\tau )b_{\stackrel{ห}{s}}^\nu (\tau )\sigma ^{\stackrel{ห}{s}}=N(\tau ,\stackrel{}{\sigma }),`$ (87)
$`N_{[z](flat)\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })`$ $``$ $`N_{(flat)\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })=\stackrel{~}{\lambda }_\mu (\tau )b_{\stackrel{ห}{r}}^\mu (\tau )b_{\stackrel{ห}{r}}^\mu (\tau )\stackrel{~}{\lambda }_{\mu \nu }(\tau )b_{\stackrel{ห}{s}}^\nu (\tau )\sigma ^{\stackrel{ห}{s}}=N_{\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma }).`$ (88)
Let us now restrict ourselves to configurations with $`ฯตp_s^2>0`$ and let us use the Wigner boost $`L^\mu {}_{\nu }{}^{}(\stackrel{}{p_s},p_s)`$ to boost to rest the variables $`b_{\stackrel{ห}{A}}^\mu `$, $`S_s^{\mu \nu }`$ of the following non-Darboux basis
$`x_s^\mu ,p_s^\mu ,b_{\stackrel{ห}{A}}^\mu ,S_s^{\mu \nu },\alpha ^i,\mathrm{\Pi }_i`$
of the Dirac brackets $`\{.,.\}^{}`$. The following new non-Darboux basis is obtained \[$`\stackrel{~}{x}_s^\mu `$ is no more a fourvector; we choose the sign $`\eta =signp_s^o`$ positive\]
$`\stackrel{~}{x}_s^\mu =x_s^\mu +{\displaystyle \frac{1}{2}}ฯต_\nu ^A(u(p_s))\eta _{AB}{\displaystyle \frac{ฯต_\rho ^B(u(p_s))}{p_{s\mu }}}S_s^{\nu \rho }=`$ (90)
$`=x_s^\mu {\displaystyle \frac{1}{\sqrt{ฯตp_s^2}(p_s^o+\eta \sqrt{p_s^2})}}[p_{s\nu }S_s^{\nu \mu }+\sqrt{ฯตp_s^2}(S_s^{o\mu }S_s^{o\nu }{\displaystyle \frac{p_{s\nu }p_s^\mu }{ฯตp_s^2}})]=`$ (91)
$`=x_s^\mu {\displaystyle \frac{1}{\sqrt{ฯตp_s^2}}}[\eta _A^\mu (\overline{S}_s^{\overline{o}A}{\displaystyle \frac{\overline{S}_s^{Ar}p_s^r}{p_s^o+\sqrt{ฯตp_s^2}}})+{\displaystyle \frac{p_s^\mu +2\sqrt{ฯตp_s^2}\eta ^{\mu o}}{\sqrt{ฯตp_s^2}(p_s^o+\sqrt{ฯตp_s^2})}}\overline{S}_s^{\overline{o}r}p_s^r],`$ (92)
(93)
$`p_s^\mu =p_s^\mu ,`$ (94)
(95)
$`\alpha ^i=\alpha ^i,`$ (96)
$`\mathrm{\Pi }_i=\mathrm{\Pi }_i,`$ (97)
(98)
$`b_{\stackrel{ห}{r}}^A=ฯต_\mu ^A(u(p_s))b_{\stackrel{ห}{r}}^\mu ,`$ (99)
$`\stackrel{~}{S}_s^{\mu \nu }=S_s^{\mu \nu }{\displaystyle \frac{1}{2}}ฯต_\rho ^A(u(p_s))\eta _{AB}({\displaystyle \frac{ฯต_\sigma ^B(u(p_s))}{p_{s\mu }}}p_s^\nu {\displaystyle \frac{ฯต_\sigma ^B(u(p_s))}{p_{s\nu }}}p_s^\mu )S_s^{\rho \sigma }=`$ (100)
$`=S_s^{\mu \nu }+{\displaystyle \frac{1}{\sqrt{ฯตp_s^2}(p_s^o+\sqrt{ฯตp_s^2})}}[p_{s\beta }(S_s^{\beta \mu }p_s^\nu S_s^{\beta \nu }p_s^\mu )+\sqrt{ฯตp_s^2}(S_s^{o\mu }p_s^\nu S_s^{o\nu }p_s^\mu )],`$ (101)
(102)
$`J_s^{\mu \nu }=\stackrel{~}{x}_s^\mu p_s^\nu \stackrel{~}{x}_s^\nu p_s^\mu +\stackrel{~}{S}_s^{\mu \nu }.`$ (103)
We have
$`\{\stackrel{~}{x}_s^\mu ,p_s^\nu \}{}_{}{}^{}=0,`$ (104)
$`\{\stackrel{~}{S}_s^{oi},b_{\stackrel{ห}{A}}^r\}{}_{}{}^{}={\displaystyle \frac{\delta ^{is}(p_s^rb_{\stackrel{ห}{A}}^sp_s^sb_{\stackrel{ห}{A}}^r)}{p_s^o+\sqrt{ฯตp_s^2}}},`$ (105)
$`\{\stackrel{~}{S}_s^{ij},b_{\stackrel{ห}{A}}^r\}{}_{}{}^{}=(\delta ^{ir}\delta ^{js}\delta ^{is}\delta ^{jr})b_{\stackrel{ห}{A}}^s,`$ (106)
$`\{\stackrel{~}{S}_s^{\mu \nu },\stackrel{~}{S}_s^{\alpha \beta }\}{}_{}{}^{}=C_{\gamma \delta }^{\mu \nu \alpha \beta }\stackrel{~}{S}_s^{\gamma \delta },`$ (107)
and we can define
$`\overline{S}_s^{AB}`$ $`=`$ $`ฯต_\mu ^A(u(p_s))ฯต_\nu ^B(u(p_s))S_s^{\mu \nu }[b_{\stackrel{ห}{r}}^A(\tau )b_\tau ^Bb_{\stackrel{ห}{r}}^B(\tau )b_\tau ^A]`$ (109)
$`{\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\left(J^\tau \sqrt{\mu ^2+\delta _{\stackrel{ห}{u}\stackrel{ห}{v}}(T^1)^{\stackrel{ห}{u}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_j}\right)(\tau ,\stackrel{}{\sigma })}`$
$``$ $`[b_{\stackrel{ห}{r}}^A(\tau )b_{\stackrel{ห}{s}}^B(\tau )b_{\stackrel{ห}{r}}^B(\tau )b_{\stackrel{ห}{s}}^A(\tau )]{\displaystyle d^3\sigma \sigma ^{\stackrel{ห}{r}}\left(J^\tau (T^1)^{\stackrel{ห}{s}l}\mathrm{\Pi }_l\right)(\tau ,\stackrel{}{\sigma })}.`$ (110)
Let us now add six more gauge-fixings by selecting the special family of spacelike hyperplanes $`\mathrm{\Sigma }_{\tau W}`$ orthogonal to $`p_s^\mu `$ (this is possible for $`ฯตp_s^2>0`$), which can be called the โWigner foliationโ of Minkowski spacetime. This can be done by requiring (only six conditions are independent)
$`T_{\stackrel{ห}{A}}^\mu (\tau )`$ $`=`$ $`b_{\stackrel{ห}{A}}^\mu (\tau )ฯต_{A=\stackrel{ห}{A}}^\mu (u(p_s))0`$ (112)
$`b_{\stackrel{ห}{A}}^A(\tau )=ฯต_\mu ^A(u(p_s))b_{\stackrel{ห}{A}}^\mu (\tau )\eta _{\stackrel{ห}{A}}^A.`$
Now the inverse tetrad $`b_{\stackrel{ห}{A}}^\mu `$ is equal to the polarization vectors $`ฯต_A^\mu (u(p_s))`$ \[see Appendix C\] and the indices โ$`\stackrel{ห}{r}`$โ are forced to coincide with the Wigner spin-1 indices โrโ, while $`\overline{o}=\tau `$ is a Lorentz-scalar index. One has
$`\overline{S}_s^{AB}`$ $``$ $`(\eta _r^A\eta _\tau ^B\eta _r^B\eta _\tau ^A)\overline{S}_s^{\tau r}`$ (113)
$``$ $`(\eta _r^A\eta _s^B\eta _r^B\eta _s^A)\overline{S}_s^{rs},`$ (114)
$`\overline{S}_s^{rs}`$ $``$ $`{\displaystyle d^3\sigma \left(J^\tau [\sigma ^r(T^1)^{sl}\mathrm{\Pi }_l\sigma ^s(T^1)^{rl}\mathrm{\Pi }_l]\right)(\tau ,\stackrel{}{\sigma })},`$ (116)
$`\overline{S}_s^{\tau r}`$ $``$ $`\overline{S}_s^{r\tau }={\displaystyle d^3\sigma \left(J^\tau \sigma ^r\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}\mathrm{\Pi }_i(T^1)^{vj}\mathrm{\Pi }_j}\right)(\tau ,\stackrel{}{\sigma })}.`$ (117)
The comparison of $`\overline{S}_s^{AB}`$ with $`\stackrel{~}{S}_s^{\mu \nu }`$ yields
$`\stackrel{~}{S}_s^{uv}`$ $`=`$ $`\delta ^{ur}\delta ^{vt}\overline{S}_s^{rt}`$ (118)
$`\stackrel{~}{S}_s^{ov}`$ $`=`$ $`{\displaystyle \frac{\delta ^{vr}\overline{S}_s^{rt}p_st}{p_s^0+\sqrt{ฯตp_s^2}}}.`$ (119)
The time constancy of $`T_{\stackrel{ห}{A}}^\mu 0`$ with respect to the Dirac Hamiltonian of Eq.(70) gives
$`{\displaystyle \frac{d}{d\tau }}[b_{\stackrel{ห}{r}}^\mu (\tau )ฯต_r^\mu (u(p_s))]`$ $`=`$ $`\{b_{\stackrel{ห}{r}}^\mu (\tau )ฯต_r^\mu (u(p_s)),H_D\}{}_{}{}^{}=`$ (120)
$`=`$ $`{\displaystyle \frac{1}{2}}\stackrel{~}{\lambda }^{\alpha \beta }(\tau )\{b_{\stackrel{ห}{r}}^\mu (\tau ),S_{s\alpha \beta }(\tau )\}{}_{}{}^{}=\stackrel{~}{\lambda }^{\mu \alpha }(\tau )b_{\stackrel{ห}{r}\alpha }(\tau )0`$ (121)
$``$ $`\stackrel{~}{\lambda }^{\mu \nu }(\tau )0,`$ (122)
so that the independent gauge-fixings contained in Eqs.(112) and the constraints $`\stackrel{~}{}^{\mu \nu }(\tau )0`$ form six pairs of second class constraints.
Besides Eqs.(63), now we have \[remember that $`\dot{x}_s^\mu (\tau )=\stackrel{~}{\lambda }^\mu (\tau )`$\]
$`l^\mu =b_\tau ^\mu =u^\mu (p_s),`$ (123)
$`z_\tau ^\mu (\tau )=\dot{x}_s^\mu (\tau )=\sqrt{g(\tau )}u^\mu (p_s)\dot{x}_{s\nu }(\tau )ฯต_r^\mu (u(p_s))ฯต_r^\nu (u(p_s)),`$ (124)
(125)
$`N(\tau )=\sqrt{g(\tau )}=[\dot{x}_{s\mu }(\tau )u^\mu (p_s))],\sqrt{\gamma }=1,`$ (126)
$`g_{\tau \tau }=\dot{x}_s^2,g_{rs}=ฯต{}_{}{}^{3}g_{rs}^{}=ฯต\delta _{rs},`$ (127)
$`g_{\tau r}=ฯต\dot{x}_{s\mu }ฯต_r^\mu (u(p_s))=ฯต\delta _{rs}N^s,N^r=\delta ^{ru}\dot{x}_{s\mu }ฯต_u^\mu (u(p_s)),`$ (128)
$`g^{\tau \tau }={\displaystyle \frac{1}{g}}={\displaystyle \frac{ฯต}{N^2}},g^{\tau r}={\displaystyle \frac{ฯต}{g}}\dot{x}_{s\mu }\delta ^{ru}ฯต_r^\mu (u(p_s))=ฯต{\displaystyle \frac{N^r}{N^2}},`$ (129)
$`g^{rs}=ฯต(\delta ^{rs}\delta ^{ru}\delta ^{sv}{\displaystyle \frac{\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))\dot{x}_{s\nu }ฯต_v^\nu (u(p_s))}{[\dot{x}_su(p_s)]^2}})=ฯต(\delta ^{rs}{\displaystyle \frac{N^rN^s}{N^2}}).`$ (130)
On the hyperplane $`\mathrm{\Sigma }_{\tau W}`$ all the degrees of freedom $`z^\mu (\tau ,\stackrel{}{\sigma })`$ are reduced to the four degrees of freedom $`\stackrel{~}{x}_s^\mu (\tau )`$, which replace $`x_s^\mu `$. The Dirac Hamiltonian is now $`H_D=\stackrel{~}{\lambda }^\mu (\tau )\stackrel{~}{}_\mu (\tau )`$ with
$`\stackrel{~}{}^\mu (\tau )=p_s^\mu `$ $``$ (131)
$``$ $`{\displaystyle }d^3\sigma (J^\tau [u^\mu (p_s)\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}\mathrm{\Pi }_i(T^1)^{vj}\mathrm{\Pi }_j}`$ (132)
$``$ $`ฯต_r^\mu (u(p_s))\mu (T^1)^{rl}\mathrm{\Pi }_l])(\tau ,\stackrel{}{\sigma })0.`$ (133)
To find the new Dirac brackets, one needs to evaluate the matrix of the old Dirac brackets of the second class constraints (without extracting the independent ones)
$`C=\left(\begin{array}{cccc}\{\stackrel{~}{}^{\alpha \beta },\stackrel{~}{}^{\gamma \delta }\}{}_{}{}^{}0& \{\stackrel{~}{}^{\alpha \beta },T_{\stackrel{ห}{B}}^\sigma \}{}_{}{}^{}=& & \\ & =\delta _{\stackrel{ห}{B}B}[\eta ^{\sigma \beta }ฯต_B^\alpha (u(p_s))\eta ^{\sigma \alpha }ฯต_B^\beta (u(p_s))]& & \\ \{T_{\stackrel{ห}{A}}^\rho ,\stackrel{~}{}^{\gamma \delta }\}{}_{}{}^{}=& \{T_{\stackrel{ห}{A}}^\rho ,T_{\stackrel{ห}{B}}^\sigma \}{}_{}{}^{}=0& & \\ =\delta _{\stackrel{ห}{A}A}[\eta ^{\rho \gamma }ฯต_A^\delta (u(p_s))\eta ^{\rho \delta }ฯต_A^\gamma (u(p_s))]& .& & \end{array}\right)`$ (138)
Since the constraints are redundant, this matrix has the following left and right null eigenvectors: $`\left(\begin{array}{c}a_{\alpha \beta }=a_{\beta \alpha }\\ 0\end{array}\right)`$ \[$`a_{\alpha \beta }`$ arbitrary\], $`\left(\begin{array}{c}0\\ ฯต_\sigma ^B(u(p_s))\end{array}\right)`$. Therefore, one has to find a left and right quasi-inverse $`\overline{C}`$, $`\overline{C}C=C\overline{C}=D`$, such that $`\overline{C}`$ and D have the same left and right null eigenvectors. One finds
$`\overline{C}`$ $`=`$ $`\left(\begin{array}{cc}0_{\gamma \delta \mu \nu }& \frac{1}{4}[\eta _{\gamma \tau }ฯต_\delta ^D(u(p_s))\eta _{\delta \tau }ฯต_\gamma ^D(u(p_s))]\\ \frac{1}{4}[\eta _{\sigma \nu }ฯต_\mu ^B(u(p_s))\eta _{\sigma \mu }ฯต_\nu ^B(u(p_s))]& 0_{\sigma \tau }^{BD}\end{array}\right)`$ (141)
$`\overline{C}C=C\overline{C}=D`$ $`=`$ $`\left(\begin{array}{cc}\frac{1}{2}(\eta _\mu ^\alpha \eta _\nu ^\beta \eta _\nu ^\alpha \eta _\mu ^\beta )& 0_\tau ^{\alpha \beta D}\\ 0_{A\mu \nu }^\rho & \frac{1}{2}(\eta _\tau ^\rho \eta _A^Dฯต^{D\rho }(u(p_s))ฯต_{A\tau }(u(p_s))\end{array}\right)`$ (145)
and the new Dirac brackets are
$`\{A,B\}^{}`$ $`=`$ $`\{A,B\}{}_{}{}^{}{\displaystyle \frac{1}{4}}[\{A,\stackrel{~}{}^{\gamma \delta }\}{}_{}{}^{}[\eta _{\gamma \tau }ฯต_\delta ^D(u(p_s))\eta _{\delta \tau }ฯต_\gamma ^D(u(p_s))]\{T_D^\tau ,B\}{}_{}{}^{}+`$ (146)
$`+`$ $`\{A,T_B^\sigma \}{}_{}{}^{}[\eta _{\sigma \nu }ฯต_\mu ^B(u(p_s))\eta _{\sigma \mu }ฯต_\nu ^B(u(p_s))]\{\stackrel{~}{}^{\mu \nu },B\}{}_{}{}^{}].`$ (147)
While the check of $`\{\stackrel{~}{}^{\alpha \beta },B\}{}_{}{}^{}=0`$ is immediate, we must use the relation $`b_{\stackrel{ห}{A}\mu }T_D^\mu ฯต^{D\rho }=T_{\stackrel{ห}{A}}^\rho `$ \[at this level we have $`T_{\stackrel{ห}{A}}^\mu =T_A^\mu `$\] to check $`\{T_A^\rho ,B\}{}_{}{}^{}=0`$.
Then, we find the following brackets for the remaining variables $`\stackrel{~}{x}_s^\mu ,p_s^\mu ,\alpha ^i,\mathrm{\Pi }_i`$
$`\{\stackrel{~}{x}_s^\mu ,p_s^\nu \}{}_{}{}^{}=\eta ^{\mu \nu },`$ (148)
$`\{\alpha ^i(\tau ,\stackrel{}{\sigma }),\mathrm{\Pi }_j(\tau ,\stackrel{}{\sigma }^{^{}})\}{}_{}{}^{}=\delta _j^i\delta ^3(\stackrel{}{\sigma }\stackrel{}{\sigma }^{^{}}),`$ (149)
and the following form of the generators of the โexternalโ Poincarรฉ group
$`p_s^\mu ,`$ (150)
$`J_s^{\mu \nu }`$ $`=`$ $`\stackrel{~}{x}_s^\mu p_s^\nu \stackrel{~}{x}_s^\nu p_s^\mu +\stackrel{~}{S}_s^{\mu \nu },`$ (153)
$`\stackrel{~}{S}_s^{oi}={\displaystyle \frac{\delta ^{ir}\overline{S}_s^{rs}p_s^s}{p_s^o+\sqrt{ฯตp_s^2}}},`$
$`\stackrel{~}{S}_s^{ij}=\delta ^{ir}\delta ^{js}\overline{S}_s^{rs}.`$
Let us come back to the four first class constraints $`\stackrel{~}{}^\mu (\tau )0`$, $`\{\stackrel{~}{}^\mu ,\stackrel{~}{}^\nu \}{}_{}{}^{}=0`$, of Eq.(77). They can be rewritten in the following form \[from Eqs.(22), (41) we have $`J^\tau =det(_r\alpha ^i)`$, $`(T^1)^{ri}=\delta ^{rs}_s\alpha ^i/det(_u\alpha ^k)`$\]
$`(\tau )`$ $`=`$ $`u^\mu (p_s)\stackrel{~}{}_\mu (\tau )=ฯต_sM_{sys}0,`$ (156)
$`M_{sys}={\displaystyle d^3\sigma (\tau ,\stackrel{}{\sigma })}=`$
$`=`$ $`{\displaystyle d^3\sigma \left(J^\tau \sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}\mathrm{\Pi }_i(T^1)^{vj}\mathrm{\Pi }_j}\right)(\tau ,\stackrel{}{\sigma })}=`$ (157)
$`=`$ $`{\displaystyle d^3\sigma \left[det(_r\alpha ^k)\sqrt{\mu ^2+\delta ^{uv}\frac{_u\alpha ^i_v\alpha ^j}{[det(_r\alpha ^k)]^2}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma })},`$ (158)
$`\stackrel{}{}_p(\tau )`$ $`\stackrel{def}{=}`$ $`\stackrel{}{P}_{sys}={\displaystyle d^3\sigma ^r(\tau ,\stackrel{}{\sigma })}=`$ (160)
$`=`$ $`{\displaystyle d^3\sigma \mu \left(J^\tau (T^1)^{rl}\mathrm{\Pi }_l\right)(\tau ,\stackrel{}{\sigma })}={\displaystyle d^3\sigma \mu \left[\delta ^{rs}_s\alpha ^i\mathrm{\Pi }_i\right](\tau ,\stackrel{}{\sigma })}0,`$ (161)
where $`M_{sys}`$ is the invariant mass of the fluid. The first one gives the mass spectrum of the isolated system, while the other three say that the total 3-momentum of the N particles on the hyperplane $`\mathrm{\Sigma }_{\tau W}`$ vanishes.
There is no more a restriction on $`p_s^\mu `$ in this special gauge, because $`u^\mu (p_s)=p_s^\mu /\sqrt{ฯตp_s^2}`$ gives the orientation of the Wigner hyperplanes containing the isolated system with respect to an arbitrary given external observer. Now the lapse and shift functions are
$`N`$ $`=`$ $`N_{[z](flat)}=N_{(flat)}=\lambda (\tau )=\dot{x}_s^\mu (\tau )u_\mu (p_s),`$ (162)
$`N_r`$ $`=`$ $`N_{[z](flat)r}=N_{(flat)r}=\lambda _r(\tau )=\dot{x}_s^\mu (\tau )ฯต_{r\mu }(u(p_s)),`$ (163)
so that the velocity of the origin of the coordinates on the Wigner hyperplane is
$$\dot{x}_s^\mu (\tau )=ฯต[\lambda (\tau )u^\mu (p_s)+\lambda _r(\tau )ฯต_r^\mu (u(p_s)),[u^2(p_s)=ฯต,ฯต_r^2(u(p_s))=ฯต].$$
(164)
The Dirac Hamiltonian is now
$$H_D=\lambda (\tau )(\tau )\stackrel{}{\lambda }(\tau )\stackrel{}{}_p(\tau ),$$
(165)
and we have $`\dot{\stackrel{~}{x}}_s^\mu =\{\stackrel{~}{x}_s^\mu ,H_D\}{}_{}{}^{}=\lambda (\tau )u^\mu (p_s)`$. Therefore, while the old $`x_s^\mu `$ had a velocity $`\dot{x}_s^\mu `$ not parallel to the normal $`l^\mu =u^\mu (p_s)`$ to the hyperplane as shown by Eqs.(164), the new $`\stackrel{~}{x}_s^\mu `$ has $`\dot{\stackrel{~}{x}}_s^\mu l^\mu `$ and no classical zitterbewegung. Moreover, we have that $`T_s=l\stackrel{~}{x}_s=lx_s`$ is the Lorentz-invariant rest frame time.
The canonical variables $`\stackrel{~}{x}_s^\mu `$, $`p_s^\mu `$, may be replaced by the canonical pairs $`ฯต_s=\sqrt{p_s^2}`$, $`T_s=p_s\stackrel{~}{x}_s/ฯต_s`$ \[to be gauge fixed with $`T_s\tau 0`$\]; $`\stackrel{}{k}_s=\stackrel{}{p}_s/ฯต_s=\stackrel{}{u}(p_s)`$, $`\stackrel{}{z}_s=ฯต_s(\stackrel{}{\stackrel{~}{x}}_s\frac{\stackrel{}{p}_s}{p_s^o}\stackrel{~}{x}_s^o)ฯต_s\stackrel{}{q}_s`$.
One obtains in this way a new kind of instant form of the dynamics, the โWigner-covariant 1-time rest-frame instant formโ with a universal breaking of Lorentz covariance. It is the special relativistic generalization of the non-relativistic separation of the center of mass from the relative motion \[$`H=\frac{\stackrel{}{P}^2}{2M}+H_{rel}`$\]. The role of the โexternalโ center of mass is taken by the Wigner hyperplane, identified by the point $`\stackrel{~}{x}_s^\mu (\tau )`$ and by its normal $`p_s^\mu `$. The invariant mass $`M_{sys}`$ of the system replaces the non-relativistic Hamiltonian $`H_{rel}`$ for the relative degrees of freedom, after the addition of the gauge-fixing $`T_s\tau 0`$ \[identifying the time parameter $`\tau `$, labelling the leaves of the foliation, with the Lorentz scalar time of the โexternalโ center of mass in the rest frame, $`T_s=p_s\stackrel{~}{x}_s/M_{sys}`$ and implying $`\lambda (\tau )=ฯต`$\]. After this gauge fixing the Dirac Hamiltonian would be pure gauge: $`H_D=\stackrel{}{\lambda }(\tau )\stackrel{}{}_p(\tau )`$. However, if we wish to reintroduce the evolution in the time $`\tau T_s`$ in this frozen phase space we must use the Hamiltonian \[in it the time evolution is generated by $`M_{sys}`$: it is like in the frozen Hamilton-Jacobi theory, in which the evolution can be reintroduced by using the energy generator of the Poincarรฉ group as Hamiltonian\]
$$H_D=M_{sys}\stackrel{}{\lambda }(\tau )\stackrel{}{}_p(\tau ).$$
(166)
The Hamilton equations for $`\alpha ^i(\tau ,\stackrel{}{\sigma })`$ in the Wigner covariant rest-frame instant form are equivalent to the hydrodynamical Euler equations:
$`_\tau \alpha ^i(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`\{\alpha ^i(\tau ,\stackrel{}{\sigma }),H_D\}=`$ (167)
$`=`$ $`\left({\displaystyle \frac{\delta ^{uv}_u\alpha ^i_v\alpha ^j\mathrm{\Pi }_j}{det(_r\alpha ^k)\sqrt{\mu ^2+\delta ^{uv}\frac{_u\alpha ^m_v\alpha ^n}{[det(_r\alpha ^k)]^2}\mathrm{\Pi }_m\mathrm{\Pi }_n}}}\right)(\tau ,\stackrel{}{\sigma })+`$ (168)
$`+`$ $`\lambda ^r(\tau )_r\alpha ^i(\tau ),`$ (169)
$`_\tau \mathrm{\Pi }_i(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`\{\mathrm{\Pi }_i(\tau ,\stackrel{}{\sigma }),H_d\}=(ijkcyclic)`$ (171)
$`=`$ $`{\displaystyle \frac{}{\sigma ^s}}[ฯต^{suv}_u\alpha ^j_v\alpha ^k\sqrt{\mu ^2+\delta ^{uv}{\displaystyle \frac{_u\alpha ^m_v\alpha ^n}{[det(_r\alpha ^k)]^2}}\mathrm{\Pi }_m\mathrm{\Pi }_n}+`$ (172)
$`+`$ $`{\displaystyle \frac{(\delta _i^m\delta _s^u\frac{\delta ^{uv}_v\alpha ^m}{det(_r\alpha ^l)}ฯต^{slt}ฯต_{ipq}_l\alpha ^p_t\alpha ^q)\frac{_u\alpha ^n}{det(_r\alpha ^l)}\mathrm{\Pi }_m\mathrm{\Pi }_n}{\sqrt{\mu ^2+\delta ^{uv}\frac{_u\alpha ^m_v\alpha ^n}{[det(_r\alpha ^k)]^2}\mathrm{\Pi }_m\mathrm{\Pi }_n}}}](\tau ,\stackrel{}{\sigma })+`$ (173)
$`+`$ $`\lambda ^r(\tau ){\displaystyle \frac{}{\sigma ^s}}[ฯต^{suv}_u\alpha ^j_v\alpha ^k{\displaystyle \frac{_r\alpha ^m}{det(_t\alpha ^l)}}\mathrm{\Pi }_m+`$ (174)
$`+`$ $`(\delta _i^m\delta _{rs}{\displaystyle \frac{_r\alpha ^m}{det(_t\alpha ^l)}}ฯต^{slt}ฯต_{ipq}_l\alpha ^p_t\alpha ^q)\mathrm{\Pi }_m](\tau ,\stackrel{}{\sigma }).`$ (175)
In this special gauge we have $`b_A^\mu L^\mu {}_{A}{}^{}(p_s,\stackrel{}{p}_s)`$ (the standard Wigner boost for timelike Poincarรฉ orbits), $`S_s^{\mu \nu }S_{sys}^{\mu \nu }`$ \[$`S_{sys}^r=ฯต^{ruv}d^3\sigma \sigma ^u\left(J^\tau (T^1)^{vs}\mathrm{\Pi }_s\right)(\tau ,\stackrel{}{\sigma })`$\], and the only remaining canonical variables are the non-covariant Newton-Wigner-like canonical โexternalโ 3-center-of-mass coordinate $`\stackrel{}{z}_s`$ (living on the Wigner hyperplanes) and $`\stackrel{}{k}_s`$. Now 3 degrees of freedom of the isolated system \[an โinternalโ center-of-mass 3-variable $`\stackrel{}{q}_{sys}`$ defined inside the Wigner hyperplane and conjugate to $`\stackrel{}{P}_{sys}`$\] become gauge variables \[the natural gauge fixing to the rest-frame condition $`\stackrel{}{P}_{sys}0`$ is $`\stackrel{}{X}_{sys}0`$, implying $`\lambda _r(\tau )=0`$, so that it coincides with the origin $`x_s^\mu (\tau )=z^\mu (\tau ,\stackrel{}{\sigma }=0)`$ of the Wigner hyperplane\]. The variable $`\stackrel{~}{x}_s^\mu `$ is playing the role of a kinematical โexternalโ center of mass for the isolated system and may be interpreted as a decoupled observer with his parametrized clock (point particle clock). All the fields living on the Wigner hyperplane are now either Lorentz scalar or with their 3-indices transformaing under Wigner rotations (induced by Lorentz transformations in Minkowski spacetime) as any Wigner spin 1 index.
## III External and internal canonical center of mass, Mollerโs center of energy and Fokker-Pryce center of inertia
Let us now consider the problem of the definition of the relativistic center of mass of a perfect fluid configuration, using the dust as an example. Let us remark that in the approach leading to the rest-frame instant form of dynamics on Wignerโs hyperplanes there is a splitting of this concept in an โexternalโ and an โinternalโ one. One can either look at the isolated system from an arbitrary Lorentz frame or put himself inside the Wigner hyperplane.
From outside one finds after the canonical reduction to Wigner hyperplane that there is an origin $`x_s^\mu (\tau )`$ for these hyperplanes (a covariant non-canonical centroid) and a non-covariant canonical coordinate $`\stackrel{~}{x}_s^\mu (\tau )`$ describing an โexternalโ decoupled point particle observer with a clock measuring the rest-frame time $`T_s`$. Associated with them there is the โexternalโ realization (153) of the Poincarรฉ group.
Instead, all the degrees of freedom of the isolated system (here the perfect fluid configuration) are described by canonical variables on the Wigner hyperplane restricted by the rest-frame condition $`\stackrel{}{P}_{sys}0`$, implying that an โinternalโ collective variable $`\stackrel{}{q}_{sys}`$ is a gauge variable and that only relative variables are physical degrees of freedom (a form of weak Mach principle).
Inside the Wigner hyperplane at $`\tau =0`$ there is another realization of the Poincarรฉ group, the โinternalโ Poincarรฉ group. Its generators are built by using the invariant mass $`M_{sys}`$ and the 3-momentum $`\stackrel{}{P}_{sys}`$, determined by the constraints (77), as the generators of the translations and by using the spin tensor $`\overline{S}_s^{AB}`$ as the generator of the Lorentz subalgebra
$`P^\tau `$ $`=`$ $`M_{sys}={\displaystyle d^3\sigma \left[J^\tau \sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma })},`$ (176)
$`P^r`$ $`=`$ $`\stackrel{}{P}_{sys}={\displaystyle d^3\sigma \left[J^\tau (T^1)^{ri}\mathrm{\Pi }_i\right](\tau ,\stackrel{}{\sigma })}0,`$ (177)
$`K^r`$ $`=`$ $`J^{\tau r}=\overline{S}_s^{\tau r}{\displaystyle d^3\sigma \sigma ^r\left[J^\tau \sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma })},`$ (178)
$`J^r`$ $`=`$ $`S_{sys}^r={\displaystyle \frac{1}{2}}ฯต^{ruv}\overline{S}_s^{uv}ฯต^{ruv}{\displaystyle d^3\sigma \sigma ^u\left[J^\tau (T^1)^{vi}\mathrm{\Pi }_i\right](\tau ,\stackrel{}{\sigma })}.`$ (179)
By using the methods of Ref. (where there is a complete discussion of many definitions of relativistic center-of-mass-like variables) we can build the three โinternalโ (that is inside the Wigner hyperplane) Wigner 3-vectors corresponding to the 3-vectors โcanonical center of massโ $`\stackrel{}{q}_{sys}`$, โMoller center of energyโ $`\stackrel{}{r}_{sys}`$ and โFokker-Pryce center of inertiaโ $`\stackrel{}{y}_{sys}`$ \[the analogous concepts for the Klein-Gordon field are in Ref. (based on Refs.), while for the relativistic N-body problem see Ref. and for the system of N charged scalar particles plus the electromagnetic field Ref.\].
The non-canonical โinternalโ Mรธller 3-center of energy and the associated spin 3-vector are
$`\stackrel{}{r}_{sys}`$ $`=`$ $`{\displaystyle \frac{\stackrel{}{K}}{P^\tau }}={\displaystyle \frac{1}{2P^\tau }}{\displaystyle d^3\sigma \stackrel{}{\sigma }\left[J^\tau \sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma })},`$ (180)
$`\stackrel{}{\mathrm{\Omega }}_{sys}`$ $`=`$ $`\stackrel{}{J}\stackrel{}{r}_{sys}\times \stackrel{}{P},`$ (185)
$`\{r_{sys}^r,P^s\}=\delta ^{rs},\{r_{sys}^r,P^\tau \}={\displaystyle \frac{P^r}{P^\tau }},`$
$`\{r_{sys}^r,r_{sys}^s\}={\displaystyle \frac{1}{(P^\tau )^2}}ฯต^{rsu}\mathrm{\Omega }_{sys}^u,`$
$`\{\mathrm{\Omega }_{sys}^r,\mathrm{\Omega }_{sys}^s\}=ฯต^{rsu}(\mathrm{\Omega }_{sys}^u{\displaystyle \frac{1}{(P^\tau )^2}}(\stackrel{}{\mathrm{\Omega }}_{sys}\stackrel{}{P})P^u),\{\mathrm{\Omega }_{sys}^r,P^\tau \}=0.`$
The canonical โinternalโ 3-center of mass $`\stackrel{}{q}_{sys}`$ \[$`\{q_{sys}^r,q_{sys}^s\}=0`$, $`\{q_{sys}^r,P^s\}=\delta ^{rs}`$, $`\{J^r,q_{sys}^s\}=ฯต^{rsu}q_{sys}^u`$\] is
$`\stackrel{}{q}_{sys}`$ $`=`$ $`\stackrel{}{r}_{sys}{\displaystyle \frac{\stackrel{}{J}\times \stackrel{}{\mathrm{\Omega }}_{sys}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2})}}=`$ (186)
$`=`$ $`{\displaystyle \frac{\stackrel{}{K}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}}}+{\displaystyle \frac{\stackrel{}{J}\times \stackrel{}{P}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2})}}+`$ (187)
$`+`$ $`{\displaystyle \frac{(\stackrel{}{K}\stackrel{}{P})\stackrel{}{P}}{P^\tau \sqrt{(P^\tau )^2\stackrel{}{P}^2}\left(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2}\right)}},`$ (189)
$`\stackrel{}{r}_{sys}for\stackrel{}{P}0;\{\stackrel{}{q}_{sys},P^\tau \}={\displaystyle \frac{\stackrel{}{P}}{P^\tau }}0,`$
$`\stackrel{}{S}_{qsys}`$ $`=`$ $`\stackrel{}{J}\stackrel{}{q}_{sys}\times \stackrel{}{P}=`$ (191)
$`=`$ $`{\displaystyle \frac{P^\tau \stackrel{}{J}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}}}+{\displaystyle \frac{\stackrel{}{K}\times \stackrel{}{P}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}}}{\displaystyle \frac{(\stackrel{}{J}\stackrel{}{P})\stackrel{}{P}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}\left(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2}\right)}}`$ (192)
$``$ $`\stackrel{}{S}_{sys},for\stackrel{}{P}0,S_{sys}^r=ฯต^{ruv}{\displaystyle d^3\sigma \sigma ^u\left(J^\tau (T^1)^{vs}\mathrm{\Pi }_s\right)(\tau ,\stackrel{}{\sigma })},`$ (194)
$`\{\stackrel{}{S}_{qsys},\stackrel{}{P}\}=\{\stackrel{}{S}_{qsys},\stackrel{}{q}_{sys}\}=0,\{S_{qsys}^r,S_{qsys}^s\}=ฯต^{rsu}S_{qsys}^u.`$
The โinternalโ non-canonical Fokker-Pryce 3-center of inertiaโ $`\stackrel{}{y}_{sys}`$ is
$`\stackrel{}{y}_{sys}`$ $`=`$ $`\stackrel{}{q}_{sys}+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{P}}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2})}}=\stackrel{}{r}_{sys}+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{P}}{P^\tau \sqrt{(P^\tau )^2\stackrel{}{P}^2}}},`$ (195)
$`\stackrel{}{q}_{sys}`$ $`=`$ $`\stackrel{}{r}_{sys}+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{P}}{P^\tau (P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2})}}={\displaystyle \frac{P^\tau \stackrel{}{r}_{sys}+\sqrt{(P^\tau )^2\stackrel{}{P}^2}\stackrel{}{y}_{sys}}{P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2}}},`$ (198)
$`\{y_{sys}^r,y_{sys}^s\}={\displaystyle \frac{1}{P^\tau \sqrt{(P^\tau )^2\stackrel{}{P}^2}}}ฯต^{rsu}\left[S_{sys}^u+{\displaystyle \frac{(\stackrel{}{S}_{sys}\stackrel{}{P})P^u}{\sqrt{(P^\tau )^2\stackrel{}{P}^2}(P^\tau +\sqrt{(P^\tau )^2\stackrel{}{P}^2})}}\right],`$
$`\stackrel{}{P}0`$ $``$ $`\stackrel{}{q}_{sys}\stackrel{}{r}_{sys}\stackrel{}{y}_{sys}.`$ (200)
The Wigner 3-vector $`\stackrel{}{q}_{sys}`$ is therefore the canonical 3-center of mass of the perfect fluid configuration \[since $`\stackrel{}{q}_{sys}\stackrel{}{r}_{sys}`$, it also describe that point $`z^\mu (\tau ,\stackrel{}{q}_{sys})=x_s^\mu (\tau )+q_{sys}^rฯต_r^\mu (u(p_s))`$ where the energy of the configuration is concentrated\].
There should exist a canonical transformation from the canonical basis $`\alpha ^i(\tau ,\stackrel{}{\sigma })`$, $`\mathrm{\Pi }_i(\tau ,\stackrel{}{\sigma })`$, to a new basis $`\stackrel{}{q}_{sys}`$, $`\stackrel{}{P}=\stackrel{}{P}_\varphi `$, $`\alpha _{rel}^i(\tau ,\stackrel{}{\sigma })`$, $`\mathrm{\Pi }_{reli}(\tau ,\stackrel{}{\sigma })`$ containing relative variables $`\alpha _{rel}^i(\tau ,\stackrel{}{\sigma })`$, $`\mathrm{\Pi }_{reli}(\tau ,\stackrel{}{\sigma })`$ with respect to the true center of mass of the perfect fluid configuration. To identify this final canonical basis one shall need the methods of Ref..
The gauge fixing $`\stackrel{}{q}_{sys}0`$ \[it implies $`\stackrel{}{\lambda }(\tau )=0`$\] forces all three internal center-of-mass variables to coincide with the origin $`x_s^\mu `$ of the Wigner hyperplane. We shall denote $`x_s^{(\stackrel{}{q}_{sys})\mu }(\tau )=x_s^\mu (0)+\tau u^\mu (p_s)`$ the origin in this gauge (it is a special centroid among the many possible ones; $`x_s^\mu (0)`$ is arbitrary).
As we shall see in the next Section, by adding the gauge fixings $`\stackrel{}{X}_{sys}=\stackrel{}{q}_{sys}0`$ one can show that the origin $`x_s^\mu (\tau )`$ becomes simultaneously the Dixon center of mass of an extended object and both the Pirani and Tulczyjew centroids (see Ref. for a review of these concepts in relation with the Papapetrou-Dixon-Souriau pole-dipole approximation of an extended body). The worldline $`x_s^{(\stackrel{}{q}_{sys})\mu }`$ is the unique center-of-mass worldline of special relativity in the sense of Refs..
With similar methods from the rest-frame instant form โexternalโ realization of the Poincarรฉ algebra of Eq. (153) with the generators $`p_s^\mu `$, $`J_s^{ij}=\stackrel{~}{x}_s^ip_s^j\stackrel{~}{x}_s^jp_s^i+\delta ^{ir}\delta ^{js}S_\varphi ^{rs}`$, $`K_s^i=J_s^{oi}=\stackrel{~}{x}_s^op_s^i\stackrel{~}{x}_s^ip_s^o\frac{\delta ^{ir}S_\varphi ^{rs}p_s^s}{p_s^o+ฯต_s}=\stackrel{~}{x}_s^op_s^i\stackrel{~}{x}_s^ip_s^o+\delta ^{ir}\frac{(\stackrel{}{S}_\varphi \times \stackrel{}{p}_s)^r}{p_s^o+ฯต_s}`$ \[for $`\stackrel{~}{x}_s^o=0`$ this is the Newton-Wigner decomposition of $`J_s^{\mu \nu }`$\] we can build three โexternalโ collective 3-positions (all located on the Wigner hyperplane): i) the โexternal canonical 3-center of mass $`\stackrel{}{Q}_s`$ connected with the โexternalโ canonical non-covariant center of mass $`\stackrel{~}{x}_s^\mu `$; ii) the โexternalโ Mรธller 3-center of energy $`\stackrel{}{R}_s`$ connected with the โexternalโ non-canonical and non-covariant Mรธller center of energy $`R_s^\mu `$; iii) the โexternalโ Fokker-Pryce 3-center of inertia connected with the โexternalโ covariant non-canonical Fokker-Price center of inertia $`Y_s^\mu `$ (when there are the gauge fixings $`\stackrel{}{\sigma }_{sys}0`$ it coincides with the origin $`x_s^\mu `$). It turns out that the Wigner hyperplane is the natural setting for the study of the Dixon multipoles of extended relativistic systems (see next Section) and for defining the canonical relative variables with respect to the center of mass.
The three โexternalโ 3-variables, the canonical $`\stackrel{}{Q}_s`$, the Mรธller $`\stackrel{}{R}_s`$ and the Fokker-Pryce $`\stackrel{}{Y}_s`$ built by using the rest-frame โexternalโ realization of the Poincarรฉ algebra are
$`\stackrel{}{R}_s`$ $`=`$ $`{\displaystyle \frac{1}{p_s^o}}\stackrel{}{K}_s=(\stackrel{}{\stackrel{~}{x}}_s{\displaystyle \frac{\stackrel{}{p}_s}{p_s^o}}\stackrel{~}{x}_s^o){\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{p_s^o(p_s^o+ฯต_s)}},`$ (201)
$`\stackrel{}{Q}_s`$ $`=`$ $`\stackrel{}{\stackrel{~}{x}}_s{\displaystyle \frac{\stackrel{}{p}_s}{p_s^o}}\stackrel{~}{x}_s^o={\displaystyle \frac{\stackrel{}{z}_s}{ฯต_s}}=\stackrel{}{R}_s+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{p_s^o(p_s^o+ฯต_s)}}={\displaystyle \frac{p_s^o\stackrel{}{R}_s+ฯต_s\stackrel{}{Y}_s}{p_s^o+ฯต_s}},`$ (202)
$`\stackrel{}{Y}_s`$ $`=`$ $`\stackrel{}{Q}_s+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{ฯต_s(p_s^o+ฯต_s)}}=\stackrel{}{R}_s+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{p_s^oฯต_s}},`$ (207)
$`\{R_s^r,R_s^s\}={\displaystyle \frac{1}{(p_s^o)^2}}ฯต^{rsu}\mathrm{\Omega }_s^u,\stackrel{}{\mathrm{\Omega }}_s=\stackrel{}{J}_s\stackrel{}{R}_s\times \stackrel{}{p}_s,`$
$`\{Y_s^r,Y_s^s\}={\displaystyle \frac{1}{ฯต_sp_s^o}}ฯต^{rsu}\left[S_{sys}^u+{\displaystyle \frac{(\stackrel{}{S}_{sys}\stackrel{}{p}_s)p_s^u}{ฯต_s(p_s^o+ฯต_s)}}\right],`$
$`\stackrel{}{p}_s\stackrel{}{Q}_s=\stackrel{}{p}_s\stackrel{}{R}_s=\stackrel{}{p}_s\stackrel{}{Y}_s=\stackrel{}{k}_s\stackrel{}{z}_s,`$
$`\stackrel{}{p}_s=0`$ $``$ $`\stackrel{}{Q}_s=\stackrel{}{Y}_s=\stackrel{}{R}_s,`$ (209)
with the same velocity and coinciding in the Lorentz rest frame where $`\stackrel{}{p}_s^\mu =ฯต_s(1;\stackrel{}{0})`$
In Ref. in a one-time framework without constraints and at a fixed time, it is shown that the 3-vector $`\stackrel{}{Y}_s`$ \[but not $`\stackrel{}{Q}_s`$ and $`\stackrel{}{R}_s`$\] satisfies the condition $`\{K_s^r,Y_s^s\}=Y_s^r\{Y_s^s,p_s^o\}`$ for being the space component of a 4-vector $`Y_s^\mu `$. In the enlarged canonical treatment including time variables, it is not clear which are the time components to be added to $`\stackrel{}{Q}_s`$, $`\stackrel{}{R}_s`$, $`\stackrel{}{Y}_s`$, to rebuild 4-dimesnional quantities $`\stackrel{~}{x}_s^\mu `$, $`R_s^\mu `$, $`Y_s^\mu `$, in an arbitrary Lorentz frame $`\mathrm{\Gamma }`$, in which the origin of the Wigner hyperplane is the 4-vector $`x_s^\mu =(x_s^o;\stackrel{}{x}_s)`$. We have
$`\stackrel{~}{x}_s^\mu (\tau )`$ $`=`$ $`(\stackrel{~}{x}_s^o(\tau );\stackrel{}{\stackrel{~}{x}}_s(\tau ))=x_s^\mu {\displaystyle \frac{1}{ฯต_s(p_s^o+ฯต_s)}}\left[p_{s\nu }S_s^{\nu \mu }+ฯต_s(S_s^{o\mu }S_s^{o\nu }{\displaystyle \frac{p_{s\nu }p_s^\mu }{ฯต_s^2}})\right],p_s^\mu ,`$ (210)
$`\stackrel{~}{x}_s^o`$ $`=`$ $`\sqrt{1+\stackrel{}{k}_s^2}(T_s+{\displaystyle \frac{\stackrel{}{k}_s\stackrel{}{z}_s}{ฯต_s}})=\sqrt{1+\stackrel{}{k}_s^2}(T_s+\stackrel{}{k}_s\stackrel{}{q}_s)x_s^0,p_s^o=ฯต_s\sqrt{1+\stackrel{}{k}_s^2},`$ (211)
$`\stackrel{}{\stackrel{~}{x}}_s`$ $`=`$ $`{\displaystyle \frac{\stackrel{}{z}_s}{ฯต_s}}+(T_s+{\displaystyle \frac{\stackrel{}{k}_s\stackrel{}{z}_s}{ฯต_s}})\stackrel{}{k}_s=\stackrel{}{q}_s+(T_s+\stackrel{}{k}_s\stackrel{}{q}_s)\stackrel{}{k}_s,\stackrel{}{p}_s=ฯต_s\stackrel{}{k}_s.`$ (212)
for the non-covariant (frame-dependent) canonical center of mass and its conjugate momentum.
Each Wigner hyperplane intersects the worldline of the arbitrary origin 4-vector $`x_s^\mu (\tau )=z^\mu (\tau ,\stackrel{}{0})`$ in $`\stackrel{}{\sigma }=0`$, the pseudo worldline of $`\stackrel{~}{x}_s^\mu (\tau )=z^\mu (\tau ,\stackrel{~}{\stackrel{}{\sigma }})`$ in some $`\stackrel{~}{\stackrel{}{\sigma }}`$ and the worldline of the Fokker-Pryce 4-vector $`Y_s^\mu (\tau )=z^\mu (\tau ,\stackrel{}{\sigma }_Y)`$ in some $`\stackrel{}{\sigma }_Y`$ \[on this worldline one can put the โinternal center of massโ with the gauge fixing $`\stackrel{}{q}_\varphi 0`$ ($`\stackrel{}{q}_\varphi \stackrel{}{r}_\varphi \stackrel{}{y}_\varphi `$ due to $`\stackrel{}{P}_\varphi 0`$)\]; one also has $`R_s^\mu =z^\mu (\tau ,\stackrel{}{\sigma }_R)`$. Since we have $`T_s=u(p_s)x_s=u(p_s)\stackrel{~}{x}_s\tau `$ on the Wigner hyperplane labelled by $`\tau `$, we require that also $`Y_s^\mu `$, $`R_s^\mu `$ have time components such that they too satisfy $`u(p_s)Y_s=u(p_s)R_s=T_s\tau `$. Therefore, it is reasonable to assume that $`\stackrel{~}{x}_s^\mu `$, $`Y_s^\mu `$ and $`R_s^\mu `$ satisfy the following equations consistently with Eqs.(185), (194) when $`T_s\tau `$ and $`\stackrel{}{q}_{sys}0`$
$`\stackrel{~}{x}_s^\mu `$ $`=`$ $`(\stackrel{~}{x}_s^o;\stackrel{}{\stackrel{~}{x}}_s)=(\stackrel{~}{x}_s^o;\stackrel{}{Q}_s+{\displaystyle \frac{\stackrel{}{p}_s}{p_s^o}}\stackrel{~}{x}_s^o)=`$ (213)
$`=`$ $`(\stackrel{~}{x}_s^o;{\displaystyle \frac{\stackrel{}{z}_s}{ฯต_s}}+(T_s+{\displaystyle \frac{\stackrel{}{k}_s\stackrel{}{z}_s}{ฯต_s}})\stackrel{}{k}_s)=x_s^{(\stackrel{}{q}_{sys})\mu }+ฯต_u^\mu (u(p_s))\stackrel{~}{\sigma }^u,`$ (214)
$`Y_s^\mu `$ $`=`$ $`(\stackrel{~}{x}_s^o;\stackrel{}{Y}_s)=`$ (215)
$`=`$ $`(\stackrel{~}{x}_s^o;{\displaystyle \frac{1}{ฯต_s}}[\stackrel{}{z}_s+{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{ฯต_s[1+u^o(p_s)]}}]+(T_s+{\displaystyle \frac{\stackrel{}{k}_s\stackrel{}{z}_s}{ฯต_s}})\stackrel{}{k}_s)=`$ (216)
$`=`$ $`\stackrel{~}{x}_s^\mu +\eta _r^\mu {\displaystyle \frac{(\stackrel{}{S}_{sys}\times \stackrel{}{p}_s)^r}{ฯต_s[1+u^o(p_s)]}}=`$ (217)
$`=`$ $`x_s^{(\stackrel{}{q}_{sys})\mu }+ฯต_u^\mu (u(p_s))\sigma _Y^u,`$ (218)
$`R_s^\mu `$ $`=`$ $`(\stackrel{~}{x}_s^o;\stackrel{}{R}_s)=`$ (219)
$`=`$ $`(\stackrel{~}{x}_s^o;{\displaystyle \frac{1}{ฯต_s}}[\stackrel{}{z}_s{\displaystyle \frac{\stackrel{}{S}_{sys}\times \stackrel{}{p}_s}{ฯต_su^o(p_s)[1+u^o(p_s)]}}]+(T_s+{\displaystyle \frac{\stackrel{}{k}_s\stackrel{}{z}_s}{ฯต_s}})\stackrel{}{k}_s)=`$ (220)
$`=`$ $`\stackrel{~}{x}_s^\mu \eta _r^\mu {\displaystyle \frac{(\stackrel{}{S}_{sys}\times \stackrel{}{p}_s)^r}{ฯต_su^o(p_s)[1+u^o(p_s)]}}=`$ (221)
$`=`$ $`x_s^{(\stackrel{}{q}_{sys})\mu }+ฯต_u^\mu (u(p_s))\sigma _R^u,`$ (222)
$`T_s`$ $`=`$ $`u(p_s)x_s^{(\stackrel{}{q}_{sys})}=u(p_s)\stackrel{~}{x}_s=u(p_s)Y_s=u(p_s)R_s,`$ (224)
$`\stackrel{~}{\sigma }^r`$ $`=`$ $`ฯต_{r\mu }(u(p_s))[x_s^{(\stackrel{}{q}_{sys})\mu }\stackrel{~}{x}_s^\mu ]={\displaystyle \frac{ฯต_{r\mu }(u(p_s))[u_\nu (p_s)S_s^{\nu \mu }+S_s^{o\mu }]}{[1+u^o(p_s)]}}=`$ (226)
$`=`$ $`S_{sys}^{\tau r}+{\displaystyle \frac{S_{sys}^{rs}p_s^s}{ฯต_s[1+u^o(p_s)]}}=ฯต_sr_\varphi ^r+{\displaystyle \frac{S_{sys}^{rs}u^s(p_s)}{1+u^o(p_s)}}`$ (227)
$``$ $`ฯต_sq_{sys}^r+{\displaystyle \frac{S_{sys}^{rs}u^s(p_s)}{1+u^o(p_s)}}{\displaystyle \frac{S_{sys}^{rs}u^s(p_s)}{1+u^o(p_s)}},`$ (228)
$`\sigma _Y^r`$ $`=`$ $`ฯต_{r\mu }(u(p_s))[x_s^{(\stackrel{}{q}_{sys})\mu }Y_s^\mu ]=\stackrel{~}{\sigma }^rฯต_{ru}(u(p_s)){\displaystyle \frac{(\stackrel{}{S}_{sys}\times \stackrel{}{p}_s)^u}{ฯต_s[1+u^o(p_s)]}}=`$ (229)
$`=`$ $`\stackrel{~}{\sigma }^r+{\displaystyle \frac{S_{sys}^{rs}u^s(p_s)}{1+u^o(p_s)}}=ฯต_sr_{sys}^rฯต_sq_{sys}^r0,`$ (230)
$`\sigma _R^r`$ $`=`$ $`ฯต_{r\mu }(u(p_s))[x_s^{(\stackrel{}{q}_{sys})\mu }R_s^\mu ]=\stackrel{~}{\sigma }^r+ฯต_{ru}(u(p_s)){\displaystyle \frac{(\stackrel{}{S}_{sys}\times \stackrel{}{p}_s)^u}{ฯต_su^o(p_s)[1+u^o(p_s)]}}=`$ (231)
$`=`$ $`\stackrel{~}{\sigma }^r{\displaystyle \frac{S_{sys}^{rs}u^s(p_s)}{u^o(p_s)[1+u^o(p_s)]}}=ฯต_sr_{sys}^r+{\displaystyle \frac{[1u^o(p_s)]S_{sys}^{rs}u^s(p_s)}{u^o(p_s)[1+u^o(p_s)]}}`$ (232)
$``$ $`{\displaystyle \frac{[1u^o(p_s)]S_{sys}^{rs}u^s(p_s)}{u^o(p_s)[1+u^o(p_s)]}},`$ (233)
$``$ $`x_s^{(\stackrel{}{q}_{sys})\mu }(\tau )=Y_s^\mu ,for\stackrel{}{q}_{sys}0,`$ (235)
namely in the gauge $`\stackrel{}{q}_{sys}0`$ the external Fokker-Pryce non-canonical center of inertia coincides with the origin $`x_s^{(\stackrel{}{q}_{sys})\mu }(\tau )`$ carrying the โinternalโ center of mass (coinciding with the โinternalโ Mรถller center of energy and with the โinternalโ Fokker-Pryce center of inertia) and also being the Pirani centroid and the Tulczyjew centroid.
Therefore, if we would find the center-of-mass canonical basis, then, in the gauge $`\stackrel{}{q}_{sys}0`$ and $`T_s\tau `$, the perfect fluid configurations would have the four-momentum density peaked on the worldline $`x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)`$; the canonical variables $`\alpha _{rel}^i(\tau ,\stackrel{}{\sigma })`$, $`\mathrm{\Pi }_{reli}(\tau ,\stackrel{}{\sigma })`$ would characterize the relative motions with respect to the โmonopoleโ configuration describing the center of mass of the fluid configuration. The โmonopoleโ configurations would be identified by the vanishing of the relative variables.
Remember that the canonical center of mass lies in between the Moller center of energy and the Fokker-Pryce center of inertia and that the non-covariance region around the Fokker-Pryce 4-vector extends to a worldtube with radius (the Moller radius) $`|\stackrel{}{S}_{sys}|/P^\tau `$.
## IV Dixonโs Multipoles in Minkowski Spacetime.
Let us now look at other properties of a perfect fluid configuration on the Wigner hyperplanes, always using the dust as an explicit example. To identify which kind of collective variables describe the center of mass of a fluid configuration let us consider it as a relativistic extended body and let us study its energy-momentum tensor and its Dixon multipoles in Minkowski spacetime.
The Euler-Lagrange equations from the action (LABEL:I10) \[(32) for the dust\] are
$`\left({\displaystyle \frac{}{z^\mu }}_{\stackrel{ห}{A}}{\displaystyle \frac{}{z_{\stackrel{ห}{A}}^\mu }}\right)(\tau ,\stackrel{}{\sigma })=\eta _{\mu \nu }_{\stackrel{ห}{A}}[\sqrt{g}T^{\stackrel{ห}{A}\stackrel{ห}{B}}[\alpha ]z_{\stackrel{ห}{B}}^\nu ](\tau ,\stackrel{}{\sigma })\stackrel{}{=}\mathrm{\hspace{0.17em}0},`$ (236)
$`\left({\displaystyle \frac{}{\varphi }}_{\stackrel{ห}{A}}{\displaystyle \frac{}{_{\stackrel{ห}{A}}\varphi }}\right)(\tau ,\stackrel{}{\sigma })\stackrel{}{=}\mathrm{\hspace{0.17em}0},`$ (237)
where we introduced the energy-momentum tensor \[with a different sign with respect to the standard convention to conform with Ref.\]
$`T^{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })[\alpha ]`$ $`=`$ $`\left[{\displaystyle \frac{2}{\sqrt{g}}}{\displaystyle \frac{\delta S}{\delta g_{\stackrel{ห}{A}\stackrel{ห}{B}}}}\right](\tau ,\stackrel{}{\sigma })=`$ (238)
$`=`$ $`\left[\rho g^{\stackrel{ห}{A}\stackrel{ห}{B}}+n{\displaystyle \frac{\rho }{n}}|_s(g^{\stackrel{ห}{A}\stackrel{ห}{B}}{\displaystyle \frac{J^{\stackrel{ห}{A}}J^{\stackrel{ห}{B}}}{g_{\stackrel{ห}{C}\stackrel{ห}{D}}J^{\stackrel{ห}{C}}J^{\stackrel{ห}{D}}}})\right](\tau ,\stackrel{}{\sigma })=`$ (239)
$`=`$ $`\left[ฯต\rho U^{\stackrel{ห}{A}}U^{\stackrel{ห}{B}}+p(g^{\stackrel{ห}{A}\stackrel{ห}{B}}ฯตU^{\stackrel{ห}{A}}U^{\stackrel{ห}{B}})\right](\tau ,\stackrel{}{\sigma })`$ (240)
$`\stackrel{dust}{=}`$ $`ฯต\mu nU^{\stackrel{ห}{A}}U^{\stackrel{ห}{B}}=ฯต{\displaystyle \frac{J^{\stackrel{ห}{A}}J^{\stackrel{ห}{B}}}{N^2\sqrt{\gamma }J^\tau }}\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{v}}^{}(T^1)^{\stackrel{ห}{u}i}(T^1)^{\stackrel{ห}{v}j}\mathrm{\Pi }_i\mathrm{\Pi }_j}.`$ (242)
When $`_{\stackrel{ห}{A}}[\sqrt{g}z_{\stackrel{ห}{B}}^\mu ]=0`$, as it happens on the Wigner hyperplanes in the gauge $`T_s\tau 0`$, $`\stackrel{}{\lambda }(\tau )=0`$, we get the conservation of the energy-momentum tensor $`T^{\stackrel{ห}{A}\stackrel{ห}{B}}`$, i.e. $`_{\stackrel{ห}{A}}T^{\stackrel{ห}{A}\stackrel{ห}{B}}\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$. Otherwise, there is compensation coming from the dynamics of the surface.
As shown in Eq.(A10) the conserved, manifestly Lorentz covariant energy-momentum tensor of the perfect fluid with equation of state $`\rho =\rho (n,s)`$ \[so that $`p=n\frac{\rho }{n}|_s\rho )`$\] is
$`T^{\mu \nu }(x)[\stackrel{~}{\alpha }]`$ $`=`$ $`\left[ฯต\rho {}_{}{}^{4}g_{}^{\mu \nu }+n{\displaystyle \frac{\rho }{n}}|_s({}_{}{}^{4}g_{}^{\mu \nu }{\displaystyle \frac{J^\mu J^\nu }{{}_{}{}^{4}g_{\alpha \beta }^{}J^\alpha J^\beta }})\right](x)=`$ (243)
$`=`$ $`\left[ฯต\rho U^\mu U^\nu +p({}_{}{}^{4}g_{}^{\mu \nu }ฯตU^\mu U^\nu )\right](x)=`$ (244)
$`=`$ $`\left[ฯต(\rho +p)U^\mu U^\nu +p{}_{}{}^{4}g_{}^{\mu \nu }\right](x)`$ (245)
$`\stackrel{dust}{=}`$ $`ฯต\mu \left[nU^\mu U^\nu \right](x),`$ (246)
$`nU^\mu `$ $`=`$ $`J^\mu =ฯต^{\mu \nu \rho \sigma }_\nu \stackrel{~}{\alpha }^1_\rho \stackrel{~}{\alpha }^2_\sigma \stackrel{~}{\alpha }^3=nz_{\stackrel{ห}{A}}^\mu U^{\stackrel{ห}{A}}=`$ (248)
$`=`$ $`z_{\stackrel{ห}{A}}^\mu J^{\stackrel{ห}{A}}=z_\tau ^\mu J^\tau +z_{\stackrel{ห}{r}}^\mu J^{\stackrel{ห}{r}}.`$ (249)
Therefore, in $`\mathrm{\Sigma }_\tau `$-adapted coordinates on each $`\mathrm{\Sigma }_\tau `$ we get
$`T^{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })[\alpha ]`$ $`=`$ $`z_\mu ^{\stackrel{ห}{A}}(\tau ,\stackrel{}{\sigma })z_\nu ^{\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })T^{\mu \nu }(x=z(\tau ,\stackrel{}{\sigma }))[\stackrel{~}{\alpha }]=`$ (250)
$`=`$ $`z_\mu ^{\stackrel{ห}{A}}(\tau ,\stackrel{}{\sigma })z_\nu ^{\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })T^{\mu \nu }(\tau ,\stackrel{}{\sigma })[\alpha =\stackrel{~}{\alpha }z],`$ (251)
On Wigner hyperplanes, where Eqs.(130) hold and where we have
$`z^\mu (\tau ,\stackrel{}{\sigma })`$ $`=`$ $`x_s^\mu (\tau )+ฯต_u^\mu (u(p_s))\sigma ^u,`$ (252)
$`{}_{}{}^{4}\eta _{}^{\mu \nu }`$ $`=`$ $`ฯต\left[u^\mu (p_s)u^\nu (p_s){\displaystyle \underset{r=1}{\overset{3}{}}}ฯต_r^\mu (u(p_s))ฯต_r^\nu (u(p_s))\right],`$ (254)
$`N`$ $`=`$ $`\dot{x}_su(p_s),N^r=\delta ^{ru}\dot{x}_{s\mu }ฯต_u^\mu (u(p_s)),`$ (256)
$`Y^r`$ $`=`$ $`{\displaystyle \frac{J^r+N^rJ^\tau }{N}}={\displaystyle \frac{J^r+\delta ^{ru}\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))J^\tau }{\dot{x}_su(p_s)}},`$ (258)
$`n`$ $`=`$ $`{\displaystyle \frac{\sqrt{ฯตg_{AB}J^AJ^B}}{N}}=\sqrt{(J^\tau )^2\delta _{rs}Y^rY^s}`$ (259)
$`\stackrel{dust}{=}`$ $`{\displaystyle \frac{\mu J^\tau }{\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}}},`$ (260)
we get \[$`\stackrel{ห}{A}=A`$\]
$`T^{\mu \nu }[x_s^\beta (\tau )+ฯต_u^\beta (u(p_s))\sigma ^u][\alpha ]=`$ (261)
$`=`$ $`[\delta _A^\tau \dot{x}_s^\mu (\tau )+\delta _A^rฯต_r^\mu (u(p_s))][\delta _B^\tau \dot{x}_s^\nu (\tau )+\delta _B^sฯต_s^\nu (u(p_s))]T^{AB}(\tau ,\stackrel{}{\sigma })=`$ (263)
$`=`$ $`\dot{x}_s^\mu (\tau )\dot{x}_s^\nu (\tau )T^{\tau \tau }(\tau ,\stackrel{}{\sigma })+ฯต_r^\mu (u(p_s))ฯต_s^\nu (u(p_s))T^{rs}(\tau ,\stackrel{}{\sigma })+`$ (264)
$`+`$ $`[\dot{x}_s^\mu (\tau )ฯต_r^\nu (u(p_s))+\dot{x}_s^\nu (\tau )ฯต_r^\mu (u(p_s))]T^{r\tau }(\tau ,\stackrel{}{\sigma }),`$ (265)
$`T^{\tau \tau }(\tau \stackrel{}{\sigma })`$ $`=`$ $`[{\displaystyle \frac{ฯต\rho }{[\dot{x}_su(p_s)]^2}}+`$ (267)
$`+`$ $`n{\displaystyle \frac{\rho }{n}}|_s({\displaystyle \frac{ฯต}{[\dot{x}_su(p_s)]^2}}{\displaystyle \frac{(J^\tau )^2}{[\dot{x}_su(p_s)]^2[(J^\tau )^2\delta _{uv}Y^uY^v]}})](\tau ,\stackrel{}{\sigma })`$ (268)
$`\stackrel{dust}{=}`$ $`ฯต\left[{\displaystyle \frac{J^\tau }{[\dot{x}_su(p_s)]^2}}\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma }),`$ (269)
$`T^{r\tau }(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`T^{\tau r}(\tau ,\stackrel{}{\sigma })=[ฯต\rho {\displaystyle \frac{\delta ^{ru}\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))}{[\dot{x}_su(p_s)]^2}}`$ (270)
$``$ $`n{\displaystyle \frac{\rho }{n}}|_s(ฯต{\displaystyle \frac{\delta ^{ru}\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))}{[\dot{x}_su(p_s)]^2}}+`$ (271)
$`+`$ $`{\displaystyle \frac{J^rJ^\tau }{[\dot{x}_su(p_s)]^2[(J^\tau )^2\delta _{uv}Y^uY^v]}})](\tau ,\stackrel{}{\sigma })`$ (272)
$`\stackrel{dust}{=}`$ $`ฯต\left[{\displaystyle \frac{J^r}{[\dot{x}_su(p_s)]^2}}\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma }),`$ (273)
$`T^{rs}(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`[ฯต\rho (\delta ^{rs}\delta ^{ru}\delta ^{sv}{\displaystyle \frac{\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))\dot{x}_{s\nu }ฯต_v^\nu (u(p_s))}{[\dot{x}_su(p_s)]^2}})`$ (274)
$``$ $`n{\displaystyle \frac{\rho }{n}}|_s(ฯต(\delta ^{rs}\delta ^{ru}\delta ^{sv}{\displaystyle \frac{\dot{x}_{s\mu }ฯต_u^\mu (u(p_s))\dot{x}_{s\nu }ฯต_v^\nu (u(p_s))}{[\dot{x}_su(p_s)]^2}})+`$ (275)
$`+`$ $`{\displaystyle \frac{J^rJ^s}{[\dot{x}_su(p_s)]^2[\dot{x}_su(p_s)]^2[(J^\tau )^2\delta _{uv}Y^uY^v]}})](\tau ,\stackrel{}{\sigma })`$ (276)
$`\stackrel{dust}{=}`$ $`ฯต\left[{\displaystyle \frac{J^rJ^s}{[\dot{x}_su(p_s)]^2J^\tau }}\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](\tau ,\stackrel{}{\sigma }).`$ (277)
Since we have
$`\dot{x}_s^\mu (\tau )`$ $`=`$ $`\lambda ^\mu (\tau )=ฯต[u^\mu (p_s)u^\nu (p_s)ฯต_r^\mu (u(p_s))ฯต_r^\nu (u(p_s))]\dot{x}_{s\nu }(\tau )=`$ (278)
$`=`$ $`ฯต\left[u^\mu (p_s)\lambda (\tau )+ฯต_r^\mu (u(p_s))\lambda _r(\tau )\right],`$ (279)
$`\dot{x}_s^2(\tau )`$ $`=`$ $`\lambda ^2(\tau )\stackrel{}{\lambda }(\tau )>0,`$ (280)
$`U_s^\mu (\tau )`$ $`=`$ $`{\displaystyle \frac{\dot{x}_s^\mu (\tau )}{\sqrt{\dot{x}_s^2(\tau )}}}=ฯต{\displaystyle \frac{\lambda (\tau )u^\mu (p_s)+\lambda _r(\tau )ฯต_r^\mu (u(p_s))}{\sqrt{\lambda ^2(\tau )\stackrel{}{\lambda }^2(\tau )}}},`$ (281)
the timelike worldline described by the origin of the Wigner hyperplane is arbitrary (i.e. gauge dependent): $`x_s^\mu (\tau )`$ may be any covariant non-canonical centroid. As already said the real โexternalโ center of mass is the canonical non-covariant $`\stackrel{~}{x}_s^\mu (T_s)=x_s^\mu (T_s)\frac{1}{ฯต_s(p_s^o+ฯต_s)}\left[p_{s\nu }S_s^{\nu \mu }+ฯต_s(S_s^{o\mu }+S_s^{o\nu }\frac{p_{s\nu }p_s^\mu }{ฯต_s^2})\right]`$: it describes a decoupled point particle observer.
In the gauge $`T_s\tau 0`$, $`\stackrel{}{X}_{sys}=\stackrel{}{q}_{sys}0`$, implying $`\lambda (\tau )=1`$, $`\stackrel{}{\lambda }(\tau )=0`$ \[$`g_{\tau \tau }=ฯต`$, $`N=1`$, $`N^r=g_{\tau r}=0`$\], we get $`\dot{x}_s^\mu (T_s)=u^\mu (p_s)`$. Therefore, in this gauge, we have the centroid
$$x_s^\mu (T_s)=x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)=x_s^\mu (0)+T_su^\mu (p_s),$$
(282)
which carries the fluid โinternalโ collective variable $`\stackrel{}{X}_{sys}=\stackrel{}{q}_{sys}0`$.
In this gauge we get the following form of the energy-momentum tensor \[$`Y^r=J^r`$\]
$`T^{\mu \nu }[x_s^{(\stackrel{}{q}_{sys})\beta }(T_s)+ฯต_u^\beta (u(p_s))\sigma ^u][\alpha ]`$ $`=`$ $`u^\mu (p_s)u^\nu (p_s)T^{\tau \tau }(T_s,\stackrel{}{\sigma })+`$ (283)
$`+`$ $`[u^\mu (p_s)ฯต_r^\nu (u(p_s))+u^\nu (p_s)ฯต_r^\mu (u(p_s))]T^{r\tau }(T_s,\stackrel{}{\sigma })+`$ (284)
$`+`$ $`ฯต_r^\mu (u(p_s))ฯต_s^\nu (u(p_s))T^{rs}(T_s,\stackrel{}{\sigma }),`$ (285)
$`T^{\tau \tau }(T_s,\stackrel{}{\sigma })`$ $`=`$ $`[ฯต\rho +`$ (287)
$`+`$ $`n{\displaystyle \frac{\rho }{n}}|_s(ฯต{\displaystyle \frac{(J^\tau )^2}{(J^\tau )^2\delta _{uv}Y^uY^v}})](T_s,\stackrel{}{\sigma })`$ (288)
$`\stackrel{dust}{=}`$ $`ฯต\left[J^\tau \sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](T_s,\stackrel{}{\sigma }),`$ (289)
$`T^{r\tau }(T_s,\stackrel{}{\sigma })`$ $`=`$ $`\left[n{\displaystyle \frac{\rho }{n}}|s{\displaystyle \frac{J^rJ^\tau }{(J^\tau )^2\delta _{uv}Y^uY^v}}\right](T_s,\stackrel{}{\sigma })`$ (290)
$`\stackrel{dust}{=}`$ $`ฯต\left[J^r\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](T_s,\stackrel{}{\sigma }),`$ (291)
$`T^{rs}(T_s,\stackrel{}{\sigma })`$ $`=`$ $`[ฯต\rho \delta ^{rs}n{\displaystyle \frac{\rho }{n}}|_s(ฯต\delta ^{rs}+`$ (292)
$`+`$ $`{\displaystyle \frac{J^rJ^s}{(J^\tau )^2\delta _{uv}Y^uY^v}})](T_s,\stackrel{}{\sigma })`$ (293)
$`\stackrel{dust}{=}`$ $`ฯต\left[{\displaystyle \frac{J^rJ^s}{J^\tau }}\sqrt{\mu ^2+\delta _{uv}(T^1)^{ui}(T^1)^{vj}\mathrm{\Pi }_i\mathrm{\Pi }_j}\right](T_s,\stackrel{}{\sigma }),`$ (296)
$`withtotal\mathrm{\hspace{0.17em}\hspace{0.17em}4}momentum`$
$`P_T^\mu [\alpha ]`$ $`=`$ $`{\displaystyle d^3\sigma T^{\mu \nu }[x_s^\mu (T_s)+ฯต_u^\mu (u(p_s))\sigma ^u][\alpha ]u_\nu (p_s)}=`$ (298)
$`=`$ $`P^\tau u^\mu (p_s)P^rฯต_r^\mu (u(p_s))P^\tau u^\mu (p_s)=`$ (299)
$`=`$ $`M_{sys}u^\mu (p_s)p_s^\mu ,`$ (302)
$`andtotalmass`$
$`M[\alpha ]`$ $`=`$ $`P_T^\mu [\alpha ]u_\mu (p_s)=P^\tau =M_{sys}.`$ (304)
The stress tensor of the perfect fluid configuration on the Wigner hyperplanes is $`T^{rs}(T_s,\stackrel{}{\sigma })`$.
We can rewrite the energy-momentum tensor in such a way that it acquires a form reminiscent of the energy-momentum tensor of an ideal relativistic fluid as seen from a local observer at rest (see the Eckart decomposition in Appendix B):
$`T^{\mu \nu }[\stackrel{~}{\alpha }]`$ $`=`$ $`[\rho [\alpha ,\mathrm{\Pi }]u^\mu (p_s)u^\nu (p_s)+`$ (305)
$`+`$ $`๐ซ[\alpha ,\mathrm{\Pi }][\eta ^{\mu \nu }u^\mu (p_s)u^\nu (p_s)]+`$ (306)
$`+`$ $`u^\mu (p_s)q^\nu [\alpha ,\mathrm{\Pi }]+u^\nu (p_s)q^\mu [\alpha ,\mathrm{\Pi }]+`$ (307)
$`+`$ $`T_{an}^{rs}[\alpha ,\mathrm{\Pi }]ฯต_r^\mu (u(p_s))ฯต_s^\nu (u(p_s))](T_s,\stackrel{}{\sigma }),`$ (308)
$`\rho [\alpha ,\mathrm{\Pi }]`$ $`=`$ $`T^{\tau \tau },`$ (310)
$`๐ซ[\alpha ,\mathrm{\Pi }]`$ $`=`$ $`{\displaystyle \frac{1}{3}}{\displaystyle \underset{u}{}}T^{uu},`$ (311)
$`q^\mu [\alpha ,\mathrm{\Pi }]`$ $`=`$ $`ฯต_r^\mu (u(p_s))T^{r\tau },`$ (312)
$`T_{an}^{rs}[\alpha ,\mathrm{\Pi }]`$ $`=`$ $`T^{rs}{\displaystyle \frac{1}{3}}\delta ^{rs}{\displaystyle \underset{u}{}}T^{uu},\delta _{uv}T_{an}^{uv}[\alpha ,\mathrm{\Pi }]=0,`$ (313)
where
i) the constant normal $`u^\mu (p_s)`$ to the Wigner hyperplanes replaces the hydrodynamic velocity field of the fluid;
ii) $`\rho [\alpha ,\mathrm{\Pi }](T_s,\stackrel{}{\sigma })`$ is the energy density;
iii) $`๐ซ[\alpha ,\mathrm{\Pi }](T_s,\stackrel{}{\sigma })`$ is the analogue of the pressure (sum of the thermodynamical pressure and of the non-equilibrium bulk stress or viscous pressure);
iv) $`q^\mu [\alpha ,\mathrm{\Pi }](T_s,\stackrel{}{\sigma })`$ is the analogue of the heat flow;
v) $`T_{an}^{rs}[\alpha ,\mathrm{\Pi }](T_s,\stackrel{}{\sigma })`$ is the shear (or anisotropic) stress tensor.
We can now study the manifestly Lorentz covariant Dixon multipoles for the perfect fluid configurationon the Wigner hyperplanes in the gauge $`\lambda (\tau )=1`$, $`\stackrel{}{\lambda }(\tau )=0`$ \[so that $`\dot{x}_s^\mu (T_s)=u^\mu (p_s)`$, $`\ddot{x}_s^\mu (T_s)=0`$, $`x_s^\mu (T_s)=x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)=x_s^\mu (0)+u^\mu (p_s)T_s`$\] with respect to the origin an arbitrary timelike worldline $`w^\mu (T_s)=z^\mu (T_s,\stackrel{}{\eta }(T_s))=x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)+ฯต_r^\mu (u(p_s))\eta ^r(T_s)`$. Since we have $`z^\mu (T_s,\stackrel{}{\sigma })=x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)+ฯต_r^\mu (u(p_s))\sigma ^r=w^\mu (T_s)+ฯต_r^\mu (u(p_s))[\sigma ^r\eta ^r(T_s)]\stackrel{def}{=}w^\mu (T_s)+\delta z^\mu (T_s,\stackrel{}{\sigma })`$ \[for $`\stackrel{}{\eta }(T_s)=0`$ we get the multipoles with respect to the origin of coordinates\], we obtain \[ $`(\mu _1..\mu _n)`$ means symmetrization, while $`[\mu _1..\mu _n]`$ means antisymmetrization; $`t_T^{\mu _1..\mu _n\mu \nu }(T_s,\stackrel{}{\eta }=0)=t_T^{\mu _1..\mu _n\mu \nu }(T_s)`$\]
$`t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s,\stackrel{}{\eta })`$ $`=`$ $`t_T^{(\mu _1\mathrm{}\mu _n)(\mu \nu )}(T_s,\stackrel{}{\eta })=`$ (314)
$`=`$ $`{\displaystyle d^3\sigma \delta z^{\mu _1}(T_s,\stackrel{}{\sigma })\mathrm{}\delta z^{\mu _n}(T_s,\stackrel{}{\sigma })T^{\mu \nu }[x_s^{(\stackrel{}{q}_{sys})\beta }(T_s)+ฯต_u^\beta (u(p_s))\sigma ^u][\alpha ]}=`$ (315)
$`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))ฯต_A^\mu (u(p_s))ฯต_B^\nu (u(p_s))I_T^{r_1..r_nAB}(T_s,\stackrel{}{\eta })=`$ (316)
$`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))[u^\mu (p_s)u^\nu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s,\stackrel{}{\eta })+`$ (317)
$`+`$ $`ฯต_r^\mu (u(p_s))ฯต_s^\nu (u(p_s))I_T^{r_1\mathrm{}r_nrs}(T_s,\stackrel{}{\eta })+`$ (318)
$`+`$ $`[u^\mu (p_s)ฯต_r^\nu (u(p_s))+u^\nu (p_s)ฯต_r^\mu (u(p_s))]I_T^{r_1\mathrm{}r_nr\tau }(T_s,\stackrel{}{\eta })],`$ (319)
$`I_T^{r_1..r_nAB}(T_s,\stackrel{}{\eta })`$ $`=`$ $`{\displaystyle d^3\sigma [\sigma ^{r_1}\eta ^{r_1}(T_s)]\mathrm{}[\sigma ^{r_n}\eta ^{r_n}(T_s)]T^{AB}(T_s,\stackrel{}{\sigma })[\alpha ]},`$ (320)
$`u_{\mu _1}(p_s)`$ $`t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s,\stackrel{}{\eta })=0,`$ (322)
$`For\stackrel{}{\eta }=0`$ $`n=0(monopole)I_T^{\tau \tau }(T_s)=P^\tau ,I_T^{r\tau }(T_s)=P^r,`$ (324)
$`t_T^{\mu _1\mathrm{}\mu _n\mu }{}_{\mu }{}^{}(T_s)`$ $`=`$ $`{\displaystyle }d^3\sigma \delta x_s^{\mu _1}(\stackrel{}{\sigma })\mathrm{}\delta x_s^{\mu _n}(\stackrel{}{\sigma })T^\mu {}_{\mu }{}^{}[x_s^{(\stackrel{}{q}_{sys})\beta }(T_s)+ฯต_u^\beta (u(p_s))\sigma ^u][\alpha ]=`$ (326)
$`\stackrel{def}{=}`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))I_T^{r_1\mathrm{}r_nA}{}_{A}{}^{}(T_s)`$ (327)
$`I_T^{r_1r_2A}{}_{A}{}^{}(T_s)`$ $`=`$ $`\stackrel{ห}{I}_T^{r_1r_2A}{}_{A}{}^{}(T_s){\displaystyle \frac{1}{3}}\delta ^{r_1r_2}\delta _{uv}I_T^{uvA}{}_{A}{}^{}(T_s)=i_T^{r_1r_2}(T_s){\displaystyle \frac{1}{2}}\delta ^{r_1r_2}\delta _{uv}i_T^{uv}(T_s),`$ (329)
$`\stackrel{ห}{I}_T^{r_1r_2A}{}_{A}{}^{}(T_s)`$ $`=`$ $`i_T^{r_1r_2}(T_s){\displaystyle \frac{1}{3}}\delta ^{r_1r_2}\delta _{uv}i_T^{uv}(T_s),\delta _{uv}\stackrel{ห}{I}_T^{uvA}{}_{A}{}^{}(T_s)=0,`$ (330)
$`i_T^{r_1r_2}(T_s)`$ $`=`$ $`I_T^{r_1r_2A}{}_{A}{}^{}(T_s)\delta ^{r_1r_2}\delta _{uv}I_T^{uvA}{}_{A}{}^{}(T_s),`$ (331)
$`\stackrel{~}{t}_T^{\mu _1\mathrm{}\mu _n}(T_s)`$ $`=`$ $`t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)u_\mu (p_s)u_\nu (p_s)=`$ (333)
$`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))I_T^{r_1..r_n\tau \tau }(T_s),`$ (334)
$`\stackrel{~}{t}_T^{\mu _1}(T_s)`$ $`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))I_T^{r_1\tau \tau }(T_s)=P^\tau ฯต_{r_1}^{\mu _1}(u(p_s))r_{sys}^{r_1},`$ (336)
$`\stackrel{~}{t}_T^{\mu _1\mu _2}(T_s)`$ $`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))ฯต_{r_2}^{\mu _2}(u(p_s))I_T^{r_1r_2\tau \tau }(T_s),`$ (338)
$`I_T^{r_1r_2\tau \tau }(T_s)`$ $`=`$ $`\widehat{I}_T^{r_1r_2\tau \tau }(T_s){\displaystyle \frac{1}{3}}\delta ^{r_1r_2}\delta _{uv}I_T^{uv\tau \tau }(T_s)=\stackrel{~}{i}_T^{r_1r_2}(T_s){\displaystyle \frac{1}{2}}\delta ^{r_1r_2}\delta _{uv}\stackrel{~}{i}_T^{uv}(T_s),`$ (340)
$`\widehat{I}_T^{r_1r_2\tau \tau }(T_s)`$ $`=`$ $`\stackrel{~}{i}_T^{r_1r_2}(T_s){\displaystyle \frac{1}{3}}\delta ^{r_1r_2}\delta _{uv}\stackrel{~}{i}_T^{uv}(T_s),\delta _{uv}\widehat{I}_T^{uv\tau \tau }(T_s)=0,`$ (341)
$`\stackrel{~}{i}_T^{r_1r_2}(T_s)`$ $`=`$ $`I_T^{r_1r_2\tau \tau }(T_s)\delta ^{r_1r_2}\delta _{uv}I_T^{uv\tau \tau }(T_s).`$ (342)
The Wigner covariant multipoles $`I_T^{r_1..r_n\tau \tau }(T_s)`$, $`I_T^{r_1..r_nrs}(T_s)`$, $`I_T^{r_1..r_nr\tau }(T_s)`$ are the mass, stress and momentum multipoles respectively.
The quantities $`\stackrel{ห}{I}_T^{r_1r_2A}{}_{A}{}^{}(T_s)`$ and $`i_T^{r_1r_2}(T_s)`$ are the traceless quadrupole moment and the inertia tensor defined by Thorne in Ref..
The quantities $`I_T^{r_1r_2\tau \tau }(T_s)`$ and $`\stackrel{~}{i}_T^{r_1r_2}(T_s)`$ are Dixonโs definitions of quadrupole moment and of tensor of inertia respectively.
Moreover, Dixonโs definition of โcenter of massโ of an extended object is $`\stackrel{~}{t}_T^{\mu _1}(T_s)=0`$ or $`I_T^{r\tau \tau }(T_s)=P^\tau r_{sys}^r=0`$: therefore the quantity $`\stackrel{}{r}_{sys}`$ defined in the previous equation is a non-canonical \[$`\{r_{sys}^r,r_{sys}^s\}=S_{sys}^{rs}`$\] candidate for the โinternalโ center of mass of the field configuration: its vanishing is a gauge fixing for $`\stackrel{}{P}0`$ and implies $`x_s^\mu (T_s)=x_s^{(\stackrel{}{q}_{sys})\mu }(T_s)=x_s^\mu (0)+u^\mu (p_s)T_s`$. As we have seen in the previous Section $`\stackrel{}{r}_{sys}`$ is the โinternalโ Mรธller 3-center of energy and we have $`\stackrel{}{r}_{sys}\stackrel{}{q}_{sys}\stackrel{}{y}_{sys}`$.
When $`I_T^{r\tau \tau }(T_s)=0`$, the equations $`0=\frac{dI_T^{r\tau \tau }(T_s)}{dT_s}=P^\tau \frac{dr_{sys}^r}{dT_s}=\stackrel{}{=}P^r`$ implies the correct momentum-velocity relation $`\frac{\stackrel{}{P}}{P^\tau }\stackrel{}{=}\frac{d\stackrel{}{r}_{sys}}{dT_s}0`$.
Then there are the related Dixon multipoles
$`p_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ $`=`$ $`t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)u_\nu (p_s)=p_T^{(\mu _1\mathrm{}\mu _n)\mu }(T_s)=`$ (343)
$`=`$ $`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))ฯต_A^\mu (u(p_s))I_T^{r_1..r_nA\tau }(T_s),`$ (344)
$`u_{\mu _1}(p_s)`$ $`p_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)=0,`$ (346)
$`n=0`$ $`p_T^\mu (T_s)=P_T^\mu [\alpha ]=ฯตฯต_A^\mu (u(p_s))P^Aฯตp_s^\mu ,`$ (348)
$`p_T^{\mu _1..\mu _n\mu }(T_s)u_\mu (p_s)`$ $`=`$ $`\stackrel{~}{t}_T^{\mu _1\mathrm{}\mu _n}(T_s)=ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))I_T^{r_1..r_n\tau \tau }(T_s).`$ (350)
The spin dipole is defined as
$`S_T^{\mu \nu }(T_s)[\alpha ]`$ $`=`$ $`2p_T^{[\mu \nu ]}(T_s)=2ฯต_r^{[\mu }(u(p_s))ฯต_A^{\nu ]}(u(p_s))I_T^{rA\tau }(T_s)=`$ (351)
$`=`$ $`S_s^{\mu \nu }=`$ (352)
$`=`$ $`ฯต_r^\mu (u(p_s))ฯต_s^\nu (u(p_s))S_{sys}^{rs}+[ฯต_r^\mu (u(p_s))u^\nu (p_s)ฯต_r^\nu (u(p_s))u^\mu (p_s)]S_{sys}^{\tau r},`$ (355)
$`u_\mu (p_s)S_T^{\mu \nu }(T_s)[\alpha ]=ฯต_r^\nu (u(p_s))S_{sys}^{\tau r}=\stackrel{~}{t}_T^\nu (T_s)=P^\tau ฯต_r^\nu (u(p_s))r_{sys}^r,`$
with $`u_\mu (p_s)S_T^{\mu \nu }(T_s)[\alpha ]=0`$ when $`\stackrel{~}{t}_T^{\mu _1}(T_s)=0`$ and this condition can be taken as a definition of center of mass equivalent to Dixonโs one. When this condition holds, the barycentric spin dipole is $`S_T^{\mu \nu }(T_s)[\alpha ]=2ฯต_r^{[\mu }(u(p_s))ฯต_s^{\nu ]}(u(p_s))I_T^{rs\tau }(T_s)`$, so that $`I_T^{[rs]\tau }(T_s)=ฯต_\mu ^r(u(p_s))ฯต_\nu ^s(u(p_s))S_T^{\mu \nu }(T_s)[\alpha ]`$.
As shown in Ref., if the fluid configuration has a compact support W on the Wigner hyperplanes $`\mathrm{\Sigma }_{W\tau }`$ and if $`f(x)`$ is a $`C^{\mathrm{}}`$ complex-valued scalar function on Minkowski spacetime with compact support \[so that its Fourier transform $`\stackrel{~}{f}(k)=d^4xf(x)e^{ikx}`$ is a slowly increasing entire analytic function on Minkowski spacetime ($`|(x^o+iy^o)^{q_o}\mathrm{}(x^3+iy^3)^{q_3}f(x^\mu +iy^\mu )|<C_{q_o\mathrm{}q_3}e^{a_o|y^o|+\mathrm{}+a_3|y^3|}`$, $`a_\mu >0`$, $`q_\mu `$ positive integers for every $`\mu `$ and $`C_{q_o\mathrm{}q_3}>0`$), whose inverse is $`f(x)=\frac{d^4k}{(2\pi )^4}\stackrel{~}{f}(k)e^{ikx}`$\], we have \[we consider $`\stackrel{}{\eta }=0`$ with $`\delta z^\mu =\delta x_s^\mu `$\]
$`<T^{\mu \nu },f>`$ $`=`$ $`{\displaystyle d^4xT^{\mu \nu }(x)f(x)}=`$ (356)
$`=`$ $`{\displaystyle ๐T_sd^3\sigma f(x_s+\delta x_s)T^{\mu \nu }[x_s(T_s)+\delta x_s(\stackrel{}{\sigma })][\alpha ]}=`$ (357)
$`=`$ $`{\displaystyle ๐T_sd^3\sigma \frac{d^4k}{(2\pi )^4}\stackrel{~}{f}(k)e^{ik[x_s(T_s)+\delta x_s(\stackrel{}{\sigma })]}T^{\mu \nu }[x_s(T_s)+\delta x_s(\stackrel{}{\sigma })][\alpha ]}=`$ (358)
$`=`$ $`{\displaystyle ๐T_s\frac{d^4k}{(2\pi )^4}\stackrel{~}{f}(k)e^{ikx_s(T_s)}d^3\sigma T^{\mu \nu }[x_s(T_s)+\delta x_s(\stackrel{}{\sigma })][\alpha ]}`$ (360)
$`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(i)^n}{n!}}[k_\mu ฯต_u^\mu (u(p_s))\sigma ^u]^n=`$
$`=`$ $`{\displaystyle ๐T_s\frac{d^4k}{(2\pi )^4}\stackrel{~}{f}(k)e^{ikx_s(T_s)}\underset{n=0}{\overset{\mathrm{}}{}}\frac{(i)^n}{n!}k_{\mu _1}\mathrm{}k_{\mu _n}t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)},`$ (361)
and, but only for $`f(x)`$ analytic in W , we get
$`<T^{\mu \nu },f>`$ $`=`$ $`{\displaystyle ๐T_s\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)\frac{^nf(x)}{x^{\mu _1}\mathrm{}x^{\mu _n}}}|_{x=x_s(T_s)},`$ (363)
$``$
$`T^{\mu \nu }(x)[\alpha ]`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n!}}{\displaystyle \frac{^n}{x^{\mu _1}\mathrm{}x^{\mu _n}}}{\displaystyle ๐T_s\delta ^4(xx_s(T_s))t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)}=`$ (364)
$`=`$ $`ฯต_A^\mu (u(p_s))ฯต_B^\nu (u(p_s))T^{AB}(T_s,\stackrel{}{\sigma })[\alpha ]=`$ (365)
$`=`$ $`ฯต_A^\mu (u(p_s))ฯต_B^\nu (u(p_s)){\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{()^n}{n!}}I_T^{r_1\mathrm{}r_nAB}(T_s,\stackrel{}{\eta }){\displaystyle \frac{^n\delta ^3(\stackrel{}{\sigma }\stackrel{}{\eta }(T_s))}{\sigma ^{r_1}\mathrm{}\sigma ^{r_n}}}|_{\stackrel{}{\eta }=0}.`$ (366)
For a non analytic $`f(x)`$ we have
$`<T^{\mu \nu },f>`$ $`=`$ $`{\displaystyle ๐T_s\underset{n=0}{\overset{N}{}}\frac{1}{n!}t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)\frac{^nf(x)}{x^{\mu _1}\mathrm{}x^{\mu _n}}}|_{x=x_s(T_s)}+`$ (367)
$`+`$ $`{\displaystyle ๐T_s\frac{d^4k}{(2\pi )^4}\stackrel{~}{f}(k)e^{ikx_s(T_s)}\underset{n=N+1}{\overset{\mathrm{}}{}}\frac{(i)^n}{n!}k_{\mu _1}\mathrm{}k_{\mu _n}t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)},`$ (368)
and, as shown in Ref., from the knowledge of the moments $`t_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ for all $`n>N`$ we can get $`T^{\mu \nu }(x)`$ and, thus, all the moments with $`nN`$.
In Appendix D other types of Dixonโs multipoles are analyzed. From this study it turns out that the multipolar expansion(366) may be rearranged with the help of the Hamilton equations implying $`_\mu T^{\mu \nu }\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$, so that for analytic fluid configurations from Eq.(D50) we get
$`T^{\mu \nu }(x)[\alpha ]`$ $`\stackrel{}{=}`$ $`ฯตu^{(\mu }(p_s)ฯต_A^{\nu )}(u(p_s)){\displaystyle ๐T_s\delta ^4(xx_s(T_s))P^A}+`$ (369)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{}{x^\rho }}{\displaystyle ๐T_s\delta ^4(xx_s(T_s))S_T^{\rho (\mu }(T_s)[\alpha ]u^{\nu )}(p_s)}+`$ (370)
$`+`$ $`{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{n!}}{\displaystyle \frac{^n}{x^{\mu _1}\mathrm{}x^{\mu _n}}}{\displaystyle ๐T_s\delta ^4(xx_s(T_s))_T^{\mu _1..\mu _n\mu \nu }(T_s)},`$ (371)
$`T^{\mu \nu }(w+\delta z)`$ $`=`$ $`ฯตu^{(\mu }(p_s)ฯต_A^{\nu )}(u(p_s))P^A\delta ^3(\stackrel{}{\sigma }\stackrel{}{\eta }(T_s))+`$ (373)
$`+`$ $`{\displaystyle \frac{1}{2}}S_T^{\rho (\mu }(T_s,\stackrel{}{\eta })[\alpha ]u^{\nu )}(p_s)ฯต_\rho ^r(u(p_s)){\displaystyle \frac{\delta ^3(\stackrel{}{\sigma }\stackrel{}{\eta }(t_s))}{\sigma ^r}}+`$ (374)
$`+`$ $`{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{()^n}{n!}}[{\displaystyle \frac{n+3}{n+1}}u^\mu (p_s)u^\nu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s,\stackrel{}{\eta })+`$ (375)
$`+`$ $`{\displaystyle \frac{1}{n}}[u^\mu (p_s)ฯต_r^\nu (u(p_s))+u^\nu (p_s)ฯต_r^\mu (u(p_s))]I_T^{r_1\mathrm{}r_nr\tau }(T_s,\stackrel{}{\eta })+`$ (376)
$`+`$ $`ฯต_{s_1}^\mu (u(p_s))ฯต_{s_2}^\nu (u(p_s))[I_T^{r_1\mathrm{}r_ns_1s_2}(T_s,\stackrel{}{\eta })`$ (377)
$``$ $`{\displaystyle \frac{n+1}{n}}(I_T^{(r_1\mathrm{}r_ns_1)s_2}(T_s,\stackrel{}{\eta })+I_T^{(r_1\mathrm{}r_ns_2)s_1}(T_s,\stackrel{}{\eta }))+`$ (378)
$`+`$ $`I_T^{(r_1\mathrm{}r_ns_1s_2)}(T_s,\stackrel{}{\eta })]],`$ (379)
where for $`n2`$ and $`\stackrel{}{\eta }=0`$ $`_T^{\mu _1..\mu _n\mu \nu }(T_s)=\frac{4(n1)}{n+1}J_T^{(\mu _1..\mu _{n1}|\mu |\mu _n)\nu }(T_s)`$, with the quantities $`J_T^{\mu _1..\mu _n\mu \nu \rho \sigma }(T_s)`$ being the Dixon $`2^{2+n}`$-pole inertial moment tensors given in Eqs.(LABEL:d7) \[the quadrupole and related inertia tensor are proportional to $`I_T^{r_1r_2\tau \tau }(T_s)`$\].
The equations $`_\mu T^{\mu \nu }\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$ imply the Papapetrou-Dixon-Souriau equations for the โpole-dipoleโ system $`P_T^\mu (T_s)`$ and $`S_T^{\mu \nu }(T_s)[\alpha ]`$ \[see Eqs.(D2) and (D48); here $`\stackrel{}{\eta }=0`$\]
$`{\displaystyle \frac{dP_T^\mu (T_s)}{dT_s}}`$ $`\stackrel{}{=}`$ $`0,`$ (380)
$`{\displaystyle \frac{dS_T^{\mu \nu }(T_s)[\alpha ]}{dT_s}}`$ $`\stackrel{}{=}`$ $`2P_T^{[\mu }(T_s)u^{\nu ]}(p_s)=2ฯตP^rฯต_r^{[\mu }(u(p_s))u^{\nu ]}(p_s)0.`$ (381)
The Cartesian Dixonโs multipoles could be re-expressed in terms of either spherical or STF (symmetric tracefree) multipoles \[both kinds of tensors are associated with the irreducible representations of the rotation group: one such multipole of order $`l`$ has exactly $`2l+1`$ independent components\].
## V Isentropic and Non-Isentropic Fluids.
Let us now consider isentropic ($`s=const.`$) perfect fluids. For them we have from Eqs.(22), (LABEL:I10) and (32) \[in general $`\mu `$ is not the chemical potential but only a parameter\]
$`n`$ $`=`$ $`{\displaystyle \frac{|J|}{N\sqrt{\gamma }}}={\displaystyle \frac{X}{\sqrt{\gamma }}},`$ (382)
$`\rho `$ $`=`$ $`\rho (n)=\rho ({\displaystyle \frac{|J|}{N\sqrt{\gamma }}})=\mu f({\displaystyle \frac{X}{\sqrt{\gamma }}}),`$ (384)
$`L`$ $`=`$ $`\mu N\sqrt{\gamma }f({\displaystyle \frac{X}{\sqrt{\gamma }}}).`$ (385)
Some possible equations of state for such fluids are (see also Appendix A): 1) $`p=0`$, dust: this implies
$`\rho (n)`$ $`=`$ $`\mu n=\mu {\displaystyle \frac{X}{\sqrt{\gamma }}},i.e.`$ (386)
$`f({\displaystyle \frac{X}{\sqrt{\gamma }}})`$ $`=`$ $`{\displaystyle \frac{X}{\sqrt{\gamma }}},{\displaystyle \frac{f(\frac{X}{\sqrt{\gamma }})}{X}}={\displaystyle \frac{1}{\sqrt{\gamma }}}.`$ (388)
2) $`p=k\rho (n)=n\frac{\rho (n)}{n}\rho (n)`$ ($`k1`$ because otherwise $`\rho =const.`$, $`\mu =0`$, $`f=Ts`$). For $`k=\frac{1}{3}`$ one has the photon gas. The previous differential equation for $`\rho (n)`$ implies
$`\rho (n)`$ $`=`$ $`(an)^{k+1}=\mu n^{k+1}=\mu ({\displaystyle \frac{X}{\sqrt{\gamma }}})^{k+1},(\mu =a^{k+1}),i.e.`$ (389)
$`f({\displaystyle \frac{X}{\sqrt{\gamma }}})`$ $`=`$ $`({\displaystyle \frac{X}{\sqrt{\gamma }}})^{k+1},{\displaystyle \frac{f(\frac{X}{\sqrt{\gamma }})}{X}}={\displaystyle \frac{k+1}{\sqrt{\gamma }}}({\displaystyle \frac{X}{\sqrt{\gamma }}})^k,`$ (391)
\[for $`k0`$ we recover 1)\].
More in general one can have $`k=k(s)`$: this is a non-isentropic perfect fluid with $`\rho =\rho (n,s)`$.
3) $`p=k\rho ^\gamma (n)=n\frac{\rho (n)}{n}\rho (n)`$ ($`\gamma 1`$) . It is an isentropic polytropic perfect fluid ($`\gamma =1+\frac{1}{n}`$). The differential equation for $`\rho (n)`$ implies \[$`a`$ is an integration constant; the chemical potential is $`\mu =\frac{\rho }{n}|_s`$\]
$`\rho (n)`$ $`=`$ $`{\displaystyle \frac{an}{[1k(an)^{\gamma 1}]^{\frac{1}{\gamma 1}}}}={\displaystyle \frac{an}{[1k(an)^{\frac{1}{n}}]^n}},i.e.`$ (392)
$`f({\displaystyle \frac{X}{\sqrt{\gamma }}})`$ $`=`$ $`{\displaystyle \frac{\frac{X}{\sqrt{\gamma }}}{[1k(a\frac{X}{\sqrt{\gamma }})^{\gamma 1}]^{\frac{1}{\gamma 1}}}}={\displaystyle \frac{\frac{X}{\sqrt{\gamma }}}{[1k(a\frac{X}{\sqrt{\gamma }})^{\frac{1}{n}}]^n}},`$ (394)
$`{\displaystyle \frac{f(\frac{X}{\sqrt{\gamma }})}{X}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{\gamma }}}[1k(a{\displaystyle \frac{X}{\sqrt{\gamma }}})^{\gamma 1}]^{\frac{\gamma }{\gamma 1}}={\displaystyle \frac{1}{\sqrt{\gamma }}}[1k(a{\displaystyle \frac{X}{\sqrt{\gamma }}})^{\frac{1}{n}}]^{(n+1)}.`$ (396)
Instead in Ref. a polytropic perfect fluid is defined by the equation of state \[see the last of Eqs.(B6); $`m`$ is a mass\]
$$\rho (n,ns)=mn+\frac{k(s)}{\gamma 1}(mn)^\gamma ,$$
(397)
and has pressure $`p=k(s)(mn)^\gamma =(\gamma 1)(\rho mn)`$ and chemical potential or specific enthalpy $`\mu ^{^{}}=mc^2+mk(s)\frac{\gamma }{\gamma 1}(mn)^{\gamma 1}`$ of Eq.(B3).
4) $`p=p(\rho )`$, barotropic perfect fluid. In the isentropic case one gets $`\rho =\rho (n)`$ by solving $`p(\rho (n))=n\frac{\rho (n)}{n}\rho (n)`$.
5)Relativistic ideal (Boltzmann) gas (this is a non-isentropic case):
$$p=nk_BTand\rho =mc^2n\mathrm{\Gamma }(\beta )p,\mu =\frac{\rho +p}{n}=mc^2\mathrm{\Gamma }(\beta ),$$
(398)
with $`\beta =\frac{mc^2}{k_BT}`$, $`\mathrm{\Gamma }(\beta )=\frac{K_3(\beta )}{K_2(\beta )}`$ ($`K_i`$ are modified Bessel functions). One gets the equation of state $`\rho =\rho (n,s)`$ by solving the differential equation
$$\frac{\rho }{n}|_s=mc^2\mathrm{\Gamma }(\frac{mc^2n}{n\frac{\rho }{n}|_s\rho }).$$
(399)
5a) Ultrarelativistic case $`\beta <<1`$ ($`mc^2<<k_BT`$): since we have $`\mathrm{\Gamma }(\beta )\frac{4}{\beta }+\frac{\beta }{2}+O(\beta ^3)`$, we get $`\rho =3nk_BT+\frac{m^2c^4n}{2k_BT}+O(k_BT\beta ^4)`$, namely $`p\frac{1}{3}\rho `$ and $`\rho \rho (n)=\mu n^{4/3}`$.
5b) Non-relativistic case $`\beta >>1`$ ($`k_BT<<mc^2`$): since we have $`\mathrm{\Gamma }(\beta )1+\frac{5}{2\beta }+O(\beta ^2)`$, we get $`\rho mc^2n+\frac{3}{2}p`$, so that we have to solve the differential equation $`3n\frac{\rho }{n}|_s5\rho +2mc^2n0`$. Its solution is $`\rho (n,s)mc^2n+k(s)n^{5/3}`$. To find $`k(s)`$ let us use the definition of temperature: $`p=nk_BT=k_B\frac{\rho }{s}|_n=k_Bn^{5/3}\frac{k(s)}{s}=n\frac{\rho }{n}|_s\rho =\frac{2}{3}n^{5/3}k(s)`$. This leads to the equation $`dlnk(s)=\frac{2ds}{3k_B}`$, whose solution is $`k(s)=he^{\frac{2s}{3k_B}}=e^{\frac{2(ss_o)}{3k_B}}`$ with $`h=e^{\frac{2s_o}{3k_B}}=const.`$. Therefore, in this case we get \[it is a polytropic like in Eq.(397) with $`\gamma =5/3`$ and $`k(s)=\frac{2}{3}m^{5/3}e^{\frac{2(ss_o)}{3k_B}}`$\]
$$\rho (n,s)mc^2n+n^{\frac{5}{3}}e^{\frac{2(ss_o)}{3k_B}}andT=\frac{1}{n}\frac{\rho }{s}|_n\frac{2}{3k_B}n^{\frac{5}{3}}e^{\frac{2(ss_o)}{3k_B}}.$$
(400)
The action for these fluids is \[$`s=s(\alpha ^i)`$\]
$$S=๐\tau d^3\sigma L(\alpha ^i(\tau ,\stackrel{}{\sigma }),z^\mu (\tau ,\stackrel{}{\sigma }))=๐\tau d^3\sigma N\sqrt{\gamma }\rho (n,s),$$
(401)
and we have as in Section II
$`J^\tau `$ $`=`$ $`ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{r}}\alpha ^1_{\stackrel{ห}{u}}\alpha ^2_{\stackrel{ห}{v}}\alpha ^3,`$ (402)
$`J^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{3}{}}}_\tau \alpha ^iฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k,`$ (404)
$`i,j,k,=\text{cyclic},`$
$`|J|`$ $`=`$ $`\sqrt{N^2(J^\tau )^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}[J^{\stackrel{ห}{r}}+N^{\stackrel{ห}{r}}J^\tau ][J^{\stackrel{ห}{v}}+N^{\stackrel{ห}{v}}J^\tau ]}=NX,`$ (405)
$`X`$ $`=`$ $`\sqrt{(J^\tau )^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{r}}Y^{\stackrel{ห}{s}}},`$ (407)
$`Y^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \frac{1}{N}}(J^{\stackrel{ห}{r}}+N^{\stackrel{ห}{r}}J^\tau ),`$ (408)
$`{\displaystyle \frac{X}{_\tau \alpha ^i}}`$ $`=`$ $`{\displaystyle \frac{Y^{\stackrel{ห}{r}}T_{\stackrel{ห}{r}i}}{NX}},`$ (411)
$`T_{\stackrel{ห}{t}i}={}_{}{}^{3}g_{\stackrel{ห}{t}\stackrel{ห}{r}}^{}ฯต^{\stackrel{ห}{r}\stackrel{ห}{u}\stackrel{ห}{v}}_{\stackrel{ห}{u}}\alpha ^j_{\stackrel{ห}{v}}\alpha ^k,(i,j,kcyclic),`$
$`{\displaystyle \frac{X}{N}}`$ $`=`$ $`{\displaystyle \frac{{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{r}}Y^{\stackrel{ห}{s}}}{NX}}={\displaystyle \frac{(J^\tau )^2X^2}{NX}},`$ (412)
$`{\displaystyle \frac{X}{N^{\stackrel{ห}{u}}}}`$ $`=`$ $`J^\tau {\displaystyle \frac{{}_{}{}^{3}g_{\stackrel{ห}{u}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{s}}}{NX}}.`$ (413)
In the cases 1), 2) and 3) \[in case 3) we rename $`\mu `$ the constant $`a`$\] the canonical momenta can be written in the form
$`\mathrm{\Pi }_i(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`{\displaystyle \frac{L(\tau ,\stackrel{}{\sigma })}{_\tau \alpha ^i(\tau ,\stackrel{}{\sigma })}}=\mu \left[{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}}{\displaystyle \frac{Y^{\stackrel{ห}{r}}T_{\stackrel{ห}{r}i}}{X}}\right](\tau ,\stackrel{}{\sigma }),`$ (414)
$``$ $`Y^{\stackrel{ห}{r}}={\displaystyle \frac{(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i}{\mu }}X({\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}})^1=Y^{\stackrel{ห}{r}}(X),`$ (416)
$``$ $`_\tau \alpha ^i={\displaystyle \frac{J^{\stackrel{ห}{r}}_{\stackrel{ห}{r}}\alpha ^i}{J^\tau }}={\displaystyle \frac{N^{\stackrel{ห}{r}}J^\tau NY^{\stackrel{ห}{r}}}{J^\tau }}_{\stackrel{ห}{r}}\alpha ^i=`$ (420)
$`=N^{\stackrel{ห}{r}}_{\stackrel{ห}{r}}\alpha ^i+N_{\stackrel{ห}{r}}\alpha ^i(T^1)^{\stackrel{ห}{r}j}\mathrm{\Pi }_j{\displaystyle \frac{X}{\mu J^\tau }}[{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}}]^1,`$
$``$
$`\rho _\mu (\tau ,\stackrel{}{\sigma })`$ $`=`$ $`{\displaystyle \frac{L(\tau ,\stackrel{}{\sigma })}{z_\tau ^\mu (\tau ,\stackrel{}{\sigma })}}=\left[l_\mu {\displaystyle \frac{L}{N}}ฯตz_{\stackrel{ห}{s}\mu }{}_{}{}^{3}g_{}^{\stackrel{ห}{s}\stackrel{ห}{r}}{\displaystyle \frac{L}{N^{\stackrel{ห}{r}}}}\right](\tau ,\stackrel{}{\sigma })=`$ (421)
$`=`$ $`\mu \sqrt{\gamma }\left[l_\mu {\displaystyle \frac{Nf(\frac{X}{\sqrt{\gamma }})}{N}}ฯตz_{\stackrel{ห}{s}\mu }{}_{}{}^{3}g_{}^{\stackrel{ห}{s}\stackrel{ห}{r}}N{\displaystyle \frac{f(\frac{X}{\sqrt{\gamma }})}{N^{\stackrel{ห}{r}}}}\right](\tau ,\stackrel{}{\sigma })=`$ (422)
$`=`$ $`[{\displaystyle \frac{\mu }{X}}(\sqrt{\gamma }Xf({\displaystyle \frac{X}{\sqrt{\gamma }}})+[(J^\tau )^2X^2]{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}})l_\mu +`$ (423)
$`+`$ $`{\displaystyle \frac{\mu }{X}}{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}}J^\tau Y^{\stackrel{ห}{r}}z_{\stackrel{ห}{r}\mu }](\tau ,\stackrel{}{\sigma })=`$ (424)
$`=`$ $`\left[{\displaystyle \frac{\mu }{X}}\left(\sqrt{\gamma }Xf({\displaystyle \frac{X}{\sqrt{\gamma }}})+[(J^\tau )^2X^2]{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}}\right)l_\mu \right](\tau ,\stackrel{}{\sigma })`$ (425)
$``$ $`ฯต\left[J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }\right](\tau ,\stackrel{}{\sigma })=`$ (426)
$`=`$ $`\left[\mu \sqrt{\gamma }G(X,J^\tau ,\sqrt{\gamma })l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }\right](\tau ,\stackrel{}{\sigma }),`$ (428)
$`with`$
$`G({\displaystyle \frac{X}{\sqrt{\gamma }}},{\displaystyle \frac{(J^\tau )^2}{\gamma }})`$ $`=`$ $`f({\displaystyle \frac{X}{\sqrt{\gamma }}})+{\displaystyle \frac{\frac{(J^\tau )^2}{\gamma }(\frac{X}{\sqrt{\gamma }})^2}{\frac{X}{\sqrt{\gamma }}}}{\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}}=f(n)+{\displaystyle \frac{\frac{(J^\tau )^2}{\gamma }n^2}{n}}{\displaystyle \frac{f(n)}{n}}.`$ (429)
To get the Hamiltonian expression of the constraints $`^\mu (\tau ,\stackrel{}{\sigma })0`$, we have to find the solution $`X`$ of the equation $`X^2+{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}Y^{\stackrel{ห}{r}}(X)Y^{\stackrel{ห}{s}}(X)=(J^\tau )^2`$ with $`Y^{\stackrel{ห}{r}}(X)`$ given by the second line of Eq.(429). This equation may be written in the following forms
$`X^2\left[\mu ^2+A^2({\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}})^2\right]=B^2,or`$ (430)
$`({\displaystyle \frac{X}{\sqrt{\gamma }}})^2`$ $`\left[1+{\displaystyle \frac{A^2}{\mu ^2+A^2}}\left(({\displaystyle \frac{f(x)}{x}}|_{x=\frac{X}{\sqrt{\gamma }}})^21\right)\right]={\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}},or`$ (431)
$`n^2`$ $`\left[1+{\displaystyle \frac{A^2}{\mu ^2+A^2}}\left(({\displaystyle \frac{f(n)}{n}})^21\right)\right]={\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}},`$ (432)
$`A^2`$ $`=`$ $`{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j,`$ (434)
$`B^2`$ $`=`$ $`\mu ^2(J^\tau )^2,`$ (435)
$``$ $`X=\sqrt{\gamma }n=\sqrt{\gamma }F({\displaystyle \frac{A^2}{\mu ^2}},{\displaystyle \frac{(J^\tau )^2}{\gamma }})=\sqrt{\gamma }\stackrel{~}{F}({\displaystyle \frac{A^2}{\mu ^2+A^2}},{\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}}),`$ (437)
$``$ $`\rho _\mu =\mu \sqrt{\gamma }\stackrel{~}{G}({\displaystyle \frac{A^2}{\mu ^2+A^2}},{\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}})l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }=`$ (439)
$`=`$ $`l_\mu +^{\stackrel{ห}{r}}z_{\stackrel{ห}{r}\mu }.`$ (440)
Therefore, all the dependence on the metric and on the Lagrangian coordinates and their momenta is concentrated in the 3 functions $`\sqrt{\gamma }`$, $`A^2/\mu ^2={}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j/\mu ^2`$, $`B^2/\mu ^2\gamma =(J^\tau )^2/\gamma `$.
Let us consider various cases.
1) $`p=0`$, dust. As in Section II the equation for $`X`$ and the constraints are
$`X^2`$ $`[\mu ^2+A^2]=B^2,`$ (441)
$`X`$ $`=`$ $`{\displaystyle \frac{B}{\sqrt{\mu ^2+A^2}}}={\displaystyle \frac{\mu |J^\tau |}{\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j}}}`$ (444)
$`_{A^20}|J^\tau |[1{\displaystyle \frac{A^2}{2\mu ^2}}+O(A^4)],`$
$`Y^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \frac{X}{\mu }}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i={\displaystyle \frac{|J^\tau |(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i}{\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j}}},`$ (447)
$``$
$`\rho _\mu `$ $`=`$ $`|J^\tau |\sqrt{\mu ^2+{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j}l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }.`$ (448)
2) $`p=k\rho `$, $`k1`$. The equation for $`X`$ is
$`X^2`$ $`[\mu ^2+{\displaystyle \frac{A^2}{(k+1)^2(\frac{X}{\sqrt{\gamma }})^{2k}}}]=B^2,or`$ (449)
$`({\displaystyle \frac{X}{\sqrt{\gamma }}})^2`$ $`[1+{\displaystyle \frac{A^2}{\mu ^2+A^2}}({\displaystyle \frac{1}{(k+1)^2(\frac{X}{\sqrt{\gamma }})^{2k}}}1)]={\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}},`$ (450)
$`Y^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \frac{\sqrt{\gamma }(\frac{X}{\sqrt{\gamma }})^{1k}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i}{\mu (k+1)}},`$ (452)
$`\rho _\mu `$ $`=`$ $`\mu \sqrt{\gamma }({\displaystyle \frac{X}{\sqrt{\gamma }}})^{k1}[(k+1){\displaystyle \frac{(J^\tau )^2}{\gamma }}k(){\displaystyle \frac{X}{\sqrt{\gamma }}}^2]l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }.`$ (454)
Let us define $`Z`$ as the deviation of $`X`$ from dust (for $`A^20`$ \[$`_\tau \alpha ^i=0`$\]: we have $`Z1`$. $`X=\frac{|B|}{\sqrt{\mu ^2+A^2}}Z`$. Then we get the following equation for $`Z`$
$`Z^2`$ $`\left[{\displaystyle \frac{\mu ^2}{\mu ^2+A^2}}+{\displaystyle \frac{A^2(\mu ^2+A^2)^{k1}\gamma ^k}{(k+1)^2B^{2k}}}Z^{2k}\right]=1,or`$ (455)
$`Z^2`$ $`[\alpha ^2+\beta _k^2Z^{2k}]=1,`$ (459)
$`\alpha ^2={\displaystyle \frac{\mu ^2}{\mu ^2+A^2}}_{A^20}\mathrm{\hspace{0.17em}1},`$
$`\beta _k^2={\displaystyle \frac{A^2(\mu ^2+A^2)^{k1}\gamma ^k}{(k+1)^2B^{2k}}}_{A^20}\mathrm{\hspace{0.17em}0}.`$
We may consider the following subcases:
2a) $`k=m1`$, with the equation
$`Z_1`$ $`=`$ $`Z^2,X={\displaystyle \frac{|B|}{\sqrt{\mu ^2+A^2}}}\sqrt{Z_1},`$ (460)
$`Z_1`$ $`[\alpha ^2+\beta _m^2Z_1^m]=1,or\beta _m^2Z_1^{1m}+\alpha ^2Z_11=0.`$ (462)
i) $`p=\rho `$ ($`k=m=1`$), with the equation
$`Z_1`$ $`=`$ $`{\displaystyle \frac{1\beta _1^2}{\alpha ^2}}={\displaystyle \frac{\gamma (\mu ^2+A^2)(4\frac{B^2}{\gamma }A^2)}{4\mu ^2B^2}},`$ (463)
$`X`$ $`=`$ $`{\displaystyle \frac{\sqrt{4\frac{B^2}{\gamma }A^2}}{2\mu }}={\displaystyle \frac{1}{2\mu }}\sqrt{4\mu ^2{\displaystyle \frac{(J^\tau )^2}{\gamma }}+{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j},forA^2<4{\displaystyle \frac{B^2}{\gamma }}.`$ (464)
ii) $`p=2\rho `$ ($`k=m=2`$), with the equation
$$\alpha ^2Z_1^2Z_1+\beta _2^2=0,Z_1=\frac{1}{2\alpha ^2}[1\pm \sqrt{14\alpha ^2\beta _2^2}],$$
(465)
iii) $`p=2\rho `$ ($`k=m=2`$), with the equation
$$\beta _2^2Z_1^3+\alpha ^2Z_11=0,$$
(466)
2b) $`k=\frac{1}{m}`$, with the equation
$`Z_2`$ $`=`$ $`Z^{\frac{2}{m}},X={\displaystyle \frac{|B|}{\sqrt{\mu ^2+A^2}}}Z_2^{\frac{m}{2}},`$ (468)
$`\alpha ^2Z_2^m+\beta _{1/m}^2Z_2^{m1}1=0,`$
i) $`p=\frac{1}{2}\rho `$ ($`k=\frac{1}{2}`$, $`m=2`$), with the equation
$`\alpha ^2Z_2^2+\beta _{1/2}^2Z_21=0,`$ (469)
$`Z_2`$ $`=`$ $`{\displaystyle \frac{1}{2\alpha ^2}}[\beta _{1/2}^2\pm \sqrt{\beta _{1/2}^4+4\alpha ^2}].`$ (470)
ii) $`p=\frac{1}{2}\rho `$ ($`k=\frac{1}{2}`$, $`m=2`$), with the equation
$$\beta _{1/2}^2(Z_2^1)^3+\alpha ^2(Z_2^1)^21=0.$$
(471)
iii) $`p=\frac{1}{3}\rho `$, photon gas ($`k=\frac{1}{3}`$, $`m=3`$), $`\beta _{1/3}^2=\frac{9A^2\gamma ^{1/3}}{16B^{2/3}(\mu ^2+A^2)^{2/3}}`$, with the equation
$`\alpha ^2Z_2^3+\beta _{1/3}^2Z_2^21=0,orZ_2^3+pZ_2^2+r=0,`$ (473)
$`withp={\displaystyle \frac{\beta _{1/3}^2}{\alpha ^2}}>0,r={\displaystyle \frac{1}{\alpha ^2}}<0,`$
$`Z_2`$ $`=`$ $`Y_2{\displaystyle \frac{1}{3}}p=Y_2{\displaystyle \frac{\beta _{1/3}^2}{3\alpha ^2}}_{A^20}\mathrm{\hspace{0.17em}1},`$ (475)
$`X`$ $`=`$ $`{\displaystyle \frac{|B|}{\sqrt{\mu ^2+A^2}}}(Y_2{\displaystyle \frac{\beta _{1/3}^2}{3\alpha ^2}})={\displaystyle \frac{|B|}{\sqrt{\mu ^2+A^2}}}(Y_2{\displaystyle \frac{3A^2(\mu ^2+A^2)^{1/3}\gamma ^{1/3}}{16B^{2/3}}})^{3/2},`$ (478)
$`Y_2^3+aY_2+b=0,`$
$`a`$ $`=`$ $`{\displaystyle \frac{1}{3}}p^2={\displaystyle \frac{\beta _{1/3}^4}{3\alpha ^4}}={\displaystyle \frac{27A^4(\mu ^2+A^2)^{2/3}\gamma ^{2/3}}{2^8B^{4/3}}}<0,`$ (479)
$`b`$ $`=`$ $`{\displaystyle \frac{1}{27}}(2p^3+27r)={\displaystyle \frac{1}{27\alpha ^2}}(2{\displaystyle \frac{\beta _{1/3}^6}{\alpha ^4}}27)={\displaystyle \frac{\mu ^2+A^2}{\mu ^2}}\left({\displaystyle \frac{27A^6\gamma }{2^{11}\mu ^4B^2}}1\right),`$ (481)
$`{\displaystyle \frac{b^2}{4}}+{\displaystyle \frac{a^3}{27}}={\displaystyle \frac{1}{427\alpha ^4}}(274{\displaystyle \frac{\beta _{1/3}^6}{\alpha ^4}})={\displaystyle \frac{(\mu ^2+A^2)^2}{4\mu ^4}}\left(1{\displaystyle \frac{27A^6\gamma }{2^{10}\mu ^4B^2}}\right)_{A^20}{\displaystyle \frac{1}{4}}>0,`$
$`Y_2`$ $`=`$ $`\left({\displaystyle \frac{b}{2}}+\sqrt{{\displaystyle \frac{b^2}{4}}+{\displaystyle \frac{a^3}{27}}}\right)^{1/3}\left({\displaystyle \frac{b}{2}}+\sqrt{{\displaystyle \frac{b^2}{4}}+{\displaystyle \frac{a^3}{27}}}\right)^{1/3}_{A^20}\mathrm{\hspace{0.17em}1},`$ (485)
$``$
$`X`$ $`=`$ $`|J^\tau |[{\displaystyle \frac{3A^2\gamma ^{1/3}}{16(J^\tau )^{2/3}}}+{\displaystyle \frac{1}{2^{1/3}}}(1{\displaystyle \frac{27A^6\gamma }{2^{11}\mu ^6(J^\tau )^2}}+\sqrt{1{\displaystyle \frac{27A^6\gamma }{2^{10}\mu ^6(J^\tau )^2}}})^{1/3}`$ (486)
$``$ $`{\displaystyle \frac{1}{2^{1/3}}}(1+{\displaystyle \frac{27A^6\gamma }{2^{11}\mu ^6(J^\tau )^2}}+\sqrt{1{\displaystyle \frac{27A^6\gamma }{2^{10}\mu ^6(J^\tau )^2}}})^{1/3}]^{3/2}_{A^20}`$ (487)
$`_{A^20}`$ $`|J^\tau |[1{\displaystyle \frac{9A^2\gamma ^{1/3}}{32(J^\tau )^{2/3}}}+O(A^4)],`$ (489)
$``$
$`\rho _\mu `$ $`=`$ $`{\displaystyle \frac{1}{3}}\mu \sqrt{\gamma }({\displaystyle \frac{X}{\sqrt{\gamma }}})^{2/3}[4{\displaystyle \frac{(J^\tau )^2}{\gamma }}({\displaystyle \frac{X}{\sqrt{\gamma }}})^2]l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }=`$ (490)
$`=`$ $`l_\mu +^rz_{r\mu }.`$ (491)
In the case of the photon gas we get a closed analytical form for the constraints.
3) $`p=k\rho ^\gamma `$, $`\gamma =1+\frac{1}{n}`$, $`\gamma 1`$ ($`n0`$), with the equation
$`X^2`$ $`\left[\mu ^2+A^2\left(1k(\mu {\displaystyle \frac{X}{\sqrt{\gamma }}})^{\gamma 1}\right)^{\frac{2\gamma }{\gamma 1}}\right]=B^2,or`$ (492)
$`X^2`$ $`\left[\mu ^2+A^2\left(1k(\mu {\displaystyle \frac{X}{\sqrt{\gamma }}})^{\frac{1}{n}}\right)^{2(n+1)}\right]=B^2,or`$ (493)
$`({\displaystyle \frac{X}{\sqrt{\gamma }}})^2`$ $`\left[1+{\displaystyle \frac{A^2}{\mu ^2+A^2}}\left([1k(\mu {\displaystyle \frac{X}{\sqrt{\gamma }}})^{\frac{1}{n}}]^{2(n+1)}1\right)\right]={\displaystyle \frac{B^2}{\gamma (\mu ^2+A^2)}},`$ (494)
$`Y^{\stackrel{ห}{r}}`$ $`=`$ $`{\displaystyle \frac{X}{\mu }}[1k({\displaystyle \frac{X}{\sqrt{\gamma }}})^{\frac{1}{n}}]^{n+1}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i,`$ (496)
$`\rho _\mu `$ $`=`$ $`\mu \sqrt{\gamma }{\displaystyle \frac{\frac{(J^\tau )^2}{\gamma }k\mu ^{\frac{1}{n}}(\frac{X}{\sqrt{\gamma }})^{2+\frac{1}{n}}}{\frac{X}{\sqrt{\gamma }}[1k(\mu \frac{X}{\sqrt{\gamma }})^{\frac{1}{n}}]^{n+1}}}l_\mu +J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_iz_{\stackrel{ห}{r}\mu }.`$ (498)
Let us define $`Z`$ as the deviation of $`X`$ from dust (for $`A^20`$ \[$`_\tau \alpha ^i=0`$\]: we have $`Z1`$, $`X=\frac{|B|}{\sqrt{\mu ^2+A^2}}Z`$). Then we get the following equation for $`Z`$
$$Z^2\left[\frac{\mu ^2}{\mu ^2+A^2}+\frac{A^2(\mu ^2+A^2)^{k1}\gamma ^k}{(k+1)^2B^{2k}}\left(1k(\frac{\mu |B|Z}{\sqrt{\gamma }\sqrt{\mu ^2+A^2}})^{\frac{1}{n}}\right)^{2(n+1)}\right]=1.$$
(499)
In conclusion, only in the cases of the dust and of the photon gas we get the closed analytic form of the constraints ( i.e. of the density of invariant mass $`(\tau ,\stackrel{}{\sigma })`$, because for the momentum density we have $`^{\stackrel{ห}{r}}=J^\tau (T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i`$ independently from the type of perfect fluid).
In all the other cases we have only an implicit form for them depending on the solution $`X`$ of Eq.(440) and numerical methods should be used.
## VI Coupling to ADM metric and tetrad gravity.
Let us now assume to have a globally hyperbolic, asymptotically flat at spatial infinity spacetime $`M^4`$ with the spacelike leaves $`\mathrm{\Sigma }_\tau `$ of the foliations associated with its 3+1 splittings, diffeomorphic to $`R^3`$ .
In $`\sigma _\tau `$-adapted coordinates $`\sigma ^A=(\sigma ^\tau =\tau ;\stackrel{}{\sigma })`$ corresponding to a holonomic basis \[$`d\sigma ^A`$, $`_A=/\sigma ^A`$\] for tensor fields we have \[$`N`$ and $`N^r`$ are the lapse and shift functions; $`{}_{}{}^{3}g_{rs}^{}`$ is the 3-metric of $`\mathrm{\Sigma }_\tau `$; $`l^A(\tau ,\stackrel{}{\sigma })`$ is the unit normal vector fields to $`\mathrm{\Sigma }_\tau `$\]
$`{}_{}{}^{4}g_{AB}^{}`$ $`=`$ $`g_{AB}=\{{}_{}{}^{4}g_{\tau \tau }^{}=ฯต(N^2{}_{}{}^{3}g_{rs}^{}N^rN^s);{}_{}{}^{4}g_{\tau r}^{}=ฯต{}_{}{}^{3}g_{rs}^{}N^s;{}_{}{}^{4}g_{rs}^{}=ฯต{}_{}{}^{3}g_{rs}^{}\}=`$ (500)
$`=`$ $`ฯตl_Al_B+\mathrm{}_{AB},`$ (501)
$`\mathrm{}_{AB}`$ $`=`$ $`{}_{}{}^{4}g_{AB}^{}ฯตl_Al_B.`$ (503)
A set of $`\mathrm{\Sigma }_\tau `$-adapted tetrad and cotetrad fields is $`(a)=(1),(2),(3)`$; $`{}_{}{}^{3}e_{(a)}^{r}`$ and $`{}_{}{}^{3}e_{r}^{(a)}={}_{}{}^{3}e_{(a)r}^{}`$ are triad and cotriad fields on $`\mathrm{\Sigma }_\tau `$\]
$`{}_{(\mathrm{\Sigma })}{}^{4}E_{(o)}^{A}`$ $`=`$ $`ฯตl^A=({\displaystyle \frac{1}{N}};{\displaystyle \frac{N^r}{N}}),`$ (504)
$`{}_{(\mathrm{\Sigma })}{}^{4}E_{(a)}^{A}`$ $`=`$ $`(0;{}_{}{}^{3}e_{(a)}^{r}),`$ (505)
$`{}_{(\mathrm{\Sigma })}{}^{4}E_{A}^{(o)}`$ $`=`$ $`l_A=(N;\stackrel{}{0}),`$ (507)
$`{}_{(\mathrm{\Sigma })}{}^{4}E_{A}^{(a)}`$ $`=`$ $`(N^{(a)}=N^r{}_{}{}^{3}e_{r}^{(a)};{}_{}{}^{3}e_{r}^{(a)}).`$ (508)
In these coordinates the energy-momentum tensor of the perfect fluid is
$$T^{AB}=ฯต(\rho +p)U^AU^B+p{}_{}{}^{4}g_{}^{AB}.$$
(509)
If we use the notation $`\mathrm{\Gamma }=ฯตl_AU^A=ฯตNU^\tau `$, we get
$`T_{AB}`$ $`=`$ $`{}_{}{}^{4}g_{AC}^{}{}_{}{}^{4}g_{BC}^{}T^{CD}=[ฯตl_Al_C+\mathrm{}_{AC}][ฯตl_Bl_D+\mathrm{}_{BD}]T^{CD}=`$ (510)
$`=`$ $`El_Al_B+j_Al_B+l_Aj_B+๐ฎ_{AB},`$ (511)
$`E`$ $`=`$ $`T^{AB}l_Al_B=ฯต[(\rho +p)\mathrm{\Gamma }^2p],`$ (513)
$`j_A`$ $`=`$ $`ฯต\mathrm{}_{AC}T^{CB}l_B=ฯต(\rho +p)\mathrm{\Gamma }\mathrm{}_{AB}U^B,`$ (514)
$`๐ฎ_{AB}`$ $`=`$ $`\mathrm{}_{AC}\mathrm{}_{BD}T^{CD}=ฯต[(\rho +p)\mathrm{}_{AC}U^C\mathrm{}_{BD}U^Dp\mathrm{}_{AB}].`$ (515)
The same decomposition can be referred to a non-holonomic basis \[$`\theta ^{\overline{A}}=\{\theta ^l=Nd\tau ;\theta ^r=d\sigma ^r+N^rd\tau \}`$, $`X_{\overline{A}}=\{X_l=N^1(_\tau N^r_r);_r\}`$; $`\overline{A}=(l;r)`$\] in which we have \[$`\overline{\mathrm{\Gamma }}=ฯต\overline{l}_{\overline{A}}\overline{U}^{\overline{A}}=ฯต\overline{U}^l=\sqrt{1{}_{}{}^{3}g_{rs}^{}\overline{U}^r\overline{U}^s}`$, so that $`\overline{U}^r`$ may be interpreted as the generalized boost velocity of $`\overline{U}^{\overline{A}}`$ with respect to $`\overline{l}^{\overline{A}}`$\]
$`{}_{}{}^{4}\overline{g}_{\overline{A}\overline{B}}^{}`$ $`=`$ $`\{{}_{}{}^{4}\overline{g}_{ll}^{}=ฯต;{}_{}{}^{4}\overline{g}_{lr}^{}=0;{}_{}{}^{4}\overline{g}_{rs}^{}={}_{}{}^{4}g_{rs}^{}=ฯต{}_{}{}^{3}g_{rs}^{}\}=`$ (516)
$`=`$ $`ฯต\overline{l}_{\overline{A}}\overline{l}_{\overline{B}}+\overline{\mathrm{}}_{\overline{A}\overline{B}},`$ (517)
$`\overline{l}^{\overline{A}}`$ $`=`$ $`(ฯต;\stackrel{}{0}),\overline{l}_{\overline{A}}=(1;\stackrel{}{0}),`$ (518)
$`\overline{\mathrm{}}_{ll}`$ $`=`$ $`\overline{\mathrm{}}_{lr}=0,\overline{\mathrm{}}_{rs}=ฯต{}_{}{}^{3}g_{rs}^{},\overline{\mathrm{}}_{\overline{A}\overline{B}}\overline{U}^{\overline{B}}=(0;ฯต{}_{}{}^{3}g_{rs}^{}),`$ (519)
$`\overline{T}_{\overline{A}\overline{B}}`$ $`=`$ $`ฯต(\rho +p)\overline{U}_{\overline{A}}\overline{U}_{\overline{B}}+p{}_{}{}^{4}\overline{g}_{\overline{A}\overline{B}}^{}=`$ (521)
$`=`$ $`\overline{E}\overline{l}_{\overline{A}}\overline{l}_{\overline{B}}+\overline{j}_{\overline{A}}\overline{l}_{\overline{B}}+\overline{l}_{\overline{A}}\overline{j}_{\overline{B}}+\overline{๐ฎ}_{\overline{A}\overline{B}},`$ (522)
$`\overline{E}`$ $`=`$ $`\overline{T}^{\overline{A}\overline{B}}\overline{l}_{\overline{A}}\overline{l}_{\overline{B}}=ฯต[(\rho +p)\overline{\mathrm{\Gamma }}^2p],`$ (524)
$`\overline{j}_{\overline{A}}`$ $`=`$ $`ฯต\overline{\mathrm{}}_{\overline{A}\overline{C}}\overline{T}^{\overline{C}\overline{B}}l_{\overline{B}}=(\overline{j}_l=0;\overline{j}_r=ฯต(\rho +p)\overline{\mathrm{\Gamma }}{}_{}{}^{3}g_{rs}^{}\overline{U}^s),`$ (525)
$`\overline{๐ฎ}_{\overline{A}\overline{B}}`$ $`=`$ $`\overline{\mathrm{}}_{\overline{A}\overline{C}}\overline{\mathrm{}}_{\overline{B}\overline{D}}\overline{T}^{\overline{C}\overline{D}}=`$ (526)
$`=`$ $`(\overline{๐ฎ}_{ll}=\overline{๐ฎ}_{lr}=0;\overline{๐ฎ}_{rs}=ฯต[p{}_{}{}^{3}g_{rs}^{}+(\rho +p){}_{}{}^{3}g_{ru}^{}\overline{U}^u{}_{}{}^{3}g_{sv}^{}\overline{U}^v)].`$ (527)
$`\overline{E}`$ and $`\overline{j}_r`$ are the energy and momentum densities determined by the Eulerian observers on $`\mathrm{\Sigma }_\tau `$, while $`\overline{๐ฎ}_{rs}`$ is called the โspatial stress tensorโ.
The non-holonomic basis is used to get the 3+1 decomposition (projection normal and parallel to $`\mathrm{\Sigma }_\tau `$) of Einsteinโs equations with matter $`{}_{}{}^{4}G_{}^{\mu \nu }\stackrel{}{=}\frac{ฯตc^3}{8\pi G}T^{\mu \nu }`$ \[when one does not has an action principle for matter, one cannot use the Hamiltonian ADM formalism\]: in this way one gets four restrictions on the Cauchy data \[ $`{}_{}{}^{4}\overline{G}_{}^{ll}\stackrel{}{=}\frac{ฯตc^3}{8\pi G}\overline{T}^{ll}`$ and $`{}_{}{}^{4}\overline{G}_{}^{lr}\stackrel{}{=}\frac{ฯตc^3}{8\pi G}\overline{T}^{lr}`$; they become the secondary first class superhamiltonian and supermomentum constraints in the ADM theory; $`k=c^3/8\pi G`$\]
$`\left[{}_{}{}^{3}R+({}_{}{}^{3}K)^2{}_{}{}^{3}K_{rs}^{}{}_{}{}^{3}K_{}^{rs}\right](\tau ,\stackrel{}{\sigma })\stackrel{}{=}{\displaystyle \frac{2}{k}}\overline{E}(\tau ,\stackrel{}{\sigma }),`$ (528)
$`({}_{}{}^{3}K_{}^{rs}{}_{}{}^{3}g_{}^{rs}{}_{}{}^{3}K)_{|s}(\tau ,\stackrel{}{\sigma })\stackrel{}{=}{\displaystyle \frac{1}{k}}\overline{J}^r(\tau ,\stackrel{}{\sigma }),`$ (529)
and the spatial Einsteinโs equations $`{}_{}{}^{4}\overline{G}_{}^{rs}\stackrel{}{=}\frac{ฯตc^3}{8\pi G}\overline{T}^{rs}`$. By introducing the extrinsic curvature $`{}_{}{}^{3}K_{rs}^{}`$, this last equations are written in a first order form \[it corresponds to the Hamilton equations of the ADM theory for $`{}_{}{}^{3}g_{rs}^{}`$ and $`{}_{}{}^{3}\stackrel{~}{\mathrm{\Pi }}_{}^{rs}=\frac{ฯตc^3}{8\pi G}\sqrt{\gamma }({}_{}{}^{3}K_{}^{rs}{}_{}{}^{3}g_{}^{rs}{}_{}{}^{3}K)`$; โ$`|`$โ denotes the covariant 3-derivative\]
$`_\tau {}_{}{}^{3}g_{rs}^{}(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`\left[N_{r|s}+N_{s|r}2N{}_{}{}^{3}K_{rs}^{}\right](\tau ,\stackrel{}{\sigma }),`$ (530)
$`_\tau {}_{}{}^{3}K_{rs}^{}(\tau ,\stackrel{}{\sigma })`$ $`\stackrel{}{=}`$ $`(N[{}_{}{}^{3}R_{rs}^{}+{}_{}{}^{3}K{}_{}{}^{3}K_{rs}^{}2{}_{}{}^{3}K_{ru}^{}{}_{}{}^{3}K_{}^{u}{}_{s}{}^{}]`$ (531)
$``$ $`N_{|s|r}+N^u{}_{|s}{}^{}{}_{}{}^{3}K_{ur}^{}+N^u{}_{|r}{}^{}{}_{}{}^{3}K_{us}^{}+N^u{}_{}{}^{3}K_{rs|u}^{})(\tau ,\stackrel{}{\sigma })`$ (532)
$``$ $`{\displaystyle \frac{1}{k}}[(\overline{๐ฎ}_{rs}+{\displaystyle \frac{1}{2}}{}_{}{}^{3}g_{rs}^{}(\overline{E}\overline{๐ฎ}^u{}_{u}{}^{})](\tau ,\stackrel{}{\sigma }).`$ (533)
The matter equations $`T^{\mu \nu }{}_{;\nu }{}^{}\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$ become a generalized continuity equation \[entropy conservation when the particle number conservation law is added to the system\] $`\overline{l}_{\overline{A}}T^{\overline{A}\overline{B}}{}_{;\overline{B}}{}^{}\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$ and generalized Euler equations $`\overline{\mathrm{}}_{\overline{A}\overline{B}}T^{\overline{B}\overline{C}}{}_{;\overline{C}}{}^{}\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$ \[$`_X`$ is the Lie derivative with respect to the vector field X\]
$`[_\tau \overline{E}+N\overline{j}^r{}_{|r}{}^{}\stackrel{}{=}[N(\overline{๐ฎ}^{rs}{}_{}{}^{3}K_{rs}^{}+\overline{E}{}_{}{}^{3}K)2\overline{j}^rN_{|r}+_\stackrel{}{N}\overline{E}](\tau ,\stackrel{}{\sigma }),`$ (534)
$`[_\tau \overline{j}_r+N\overline{๐ฎ}^{rs}{}_{|s}{}^{}](\tau ,\stackrel{}{\sigma })\stackrel{}{=}[N(2{}_{}{}^{3}K_{}^{rs}\overline{j}_s+\overline{j}^r{}_{}{}^{3}K)\overline{๐ฎ}^{rs}N_{|s}\overline{E}N^{|r}+_\stackrel{}{N}\overline{j}^r](\tau ,\stackrel{}{\sigma }).`$ (535)
The equation for $`\overline{๐ฎ}_{rs}`$ would follow from an equation of state or dynamical equation of the sources \[for perfect fluids it is the particle number conservation\].
This formulation is the starting point of many approaches to the post-Newtonian approximation (see for instance Refs.) and to numerical gravity (see for instance Refs.).
Instead with the action principle for the perfect fluid described with Lagrangian coordinates (containing the information on the equation of state and on the particle number and entropy conservations) coupled to the action for tetrad gravity of Ref. (it is the ADM action of metric gravity re-expressed in terms of a new parametrization of tetrad fields) we get
$`S`$ $`=`$ $`ฯตk{\displaystyle }d\tau d^3\sigma \{N{}_{}{}^{3}eฯต_{(a)(b)(c)}{}_{}{}^{3}e_{(a)}^{r}{}_{}{}^{3}e_{(b)}^{s}{}_{}{}^{3}\mathrm{\Omega }_{rs(c)}^{}+`$ (536)
$`+`$ $`{\displaystyle \frac{{}_{}{}^{3}e}{2N}}({}_{}{}^{3}G_{o}^{1})_{(a)(b)(c)(d)}{}_{}{}^{3}e_{(b)}^{r}(N_{(a)|r}_\tau {}_{}{}^{3}e_{(a)r}^{}){}_{}{}^{3}e_{(d)}^{s}(N_{(c)|s}_\tau {}_{}{}^{3}e_{(c)s}^{})\}(\tau ,\stackrel{}{\sigma })+`$ (537)
$``$ $`{\displaystyle ๐\tau d^3\sigma \{N\sqrt{\gamma }\rho (\frac{|J|}{N\sqrt{\gamma }},s)\}(\tau ,\stackrel{}{\sigma })}.`$ (538)
The superhamiltonian and supermomentum constraints of Ref. are modified in the following way by the presence of the perfect fluid
$``$ $`=`$ $`_o+0,`$ (539)
$`_r`$ $``$ $`\mathrm{\Theta }_r=\mathrm{\Theta }_{or}+_r0,`$ (540)
with $`_r=J^\tau (T^1)^{ri}\mathrm{\Pi }_i=_r\alpha ^i\mathrm{\Pi }_i`$ and with $``$ given by Eq.(50) for the dust and by Eq.(491) for the photon gas.
In the case of dust the explicit Hamiltonian form of the energy and momentum densities is
$`(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`J^\tau (\tau ,\stackrel{}{\sigma })\sqrt{\mu ^2+[{}_{}{}^{3}g_{rs}^{}(T^1(\alpha ))^{ri}(T^1(\alpha ))^{sj}\mathrm{\Pi }_i\mathrm{\Pi }_j](\tau ,\stackrel{}{\sigma })}=`$ (541)
$`=`$ $`det(_r\alpha ^i(\tau ,\stackrel{}{\sigma }))\sqrt{\mu ^2+{}_{}{}^{3}g_{}^{uv}{\displaystyle \frac{_u\alpha ^m_v\alpha ^n}{[det(_r\alpha ^k)]^2}}\mathrm{\Pi }_m\mathrm{\Pi }_n](\tau ,\stackrel{}{\sigma })},`$ (542)
$`_r(\tau ,\stackrel{}{\sigma })`$ $`=`$ $`{}_{}{}^{3}g_{rs}^{}J^\tau (\tau ,\stackrel{}{\sigma })[(T^1(\alpha ))^{si}\mathrm{\Pi }_i](\tau ,\stackrel{}{\sigma })=_r\alpha ^i(\tau ,\stackrel{}{\sigma })\mathrm{\Pi }_i(\tau ,\stackrel{}{\sigma }).`$ (544)
In any case, all the dependence of $``$ and $`^r`$ on the metric and on the Lagrangian coordinates and their momenta is concentrated in the 3 functions $`\sqrt{\gamma }`$, $`A^2/\mu ^2={}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(T^1)^{\stackrel{ห}{r}i}\mathrm{\Pi }_i(T^1)^{\stackrel{ห}{s}j}\mathrm{\Pi }_j/\mu ^2=\frac{{}_{}{}^{3}g_{}^{uv}_u\alpha ^i_v\alpha ^j\mathrm{\Pi }_i\mathrm{\Pi }J}{\mu ^2[det(_u\alpha ^k)]^2}`$, $`B^2/\mu ^2\gamma =(J^\tau )^2/\gamma =\frac{1}{\gamma }[det(_r\alpha ^i)]^2`$.
The study of the canonical reduction to the 3-orthogonal gauges will be done in a future paper.
## VII Non-Dissipative Elastic Materials.
With the same formalism we may describe relativistic continuum mechanics \[any relativistic material (non-homogeneous, pre-stressed,โฆ) in the non-dissipative regime\] and in particular a relativistic elastic continuum \[see also Refs. and their bibliography\] in the rest-frame instant form of dynamics.
Now the scalar fields $`\stackrel{~}{\alpha }^i(z(\tau ,\stackrel{}{\sigma }))=\alpha ^i(\tau ,\stackrel{}{\sigma })`$ describe the idealized โmoleculesโ of the material in an abstract 3-dimensional manifold called the โmaterial spaceโ, while $`J^{\stackrel{ห}{A}}(\tau ,\stackrel{}{\sigma })=[N\sqrt{\gamma }nU^{\stackrel{ห}{A}}](\tau ,\stackrel{}{\sigma })`$ is the matter number current with future-oriented timelike 4-velocity vector field $`U^{\stackrel{ห}{A}}(\tau ,\stackrel{}{\sigma })`$; $`n`$ is a scalar field describing the local rest-frame matter number density. The quantity $`_{\stackrel{ห}{A}}\alpha ^i(\tau ,\stackrel{}{\sigma })=z_{\stackrel{ห}{A}}^\mu (\tau ,\stackrel{}{\sigma })_\mu \stackrel{~}{\alpha }^i(z)`$ is called the โrelativistic deformation gradientโ in $`\mathrm{\Sigma }_\tau `$-adapted coordinates.
The material space inherits a Riemannian (symmetric and positive definite) 3-metric from the spacetime $`M^4`$
$$G^{ij}={}_{}{}^{4}g_{}^{\mu \nu }_\mu \stackrel{~}{\alpha }^i_\nu \stackrel{~}{\alpha }^j={}_{}{}^{4}g_{}^{\stackrel{ห}{A}\stackrel{ห}{B}}_{\stackrel{ห}{A}}\alpha ^i_{\stackrel{ห}{B}}\alpha ^j.$$
(545)
Its inverse $`G_{ij}`$ carries the information about the actual distances of adjacent molecules in the local rest frame.
For an ideal fluid the 3-form $`\eta `$ of Eq(6) gives the volume element in the material space, which is sufficient to describe the mechanical properties of an ideal fluid. Since we have that $`n=\frac{|J|}{N\sqrt{\gamma }}`$ is a scalar, we can evaluate $`n`$ in the local rest frame at $`z`$ where $`U^\mu (z)=(\frac{1}{\sqrt{ฯต{}_{}{}^{4}\overline{g}_{\overline{o}\overline{o}}^{}}};\stackrel{}{0})`$ and $`_o\stackrel{~}{\alpha }^i|_z=0`$ \[in the local adapted non-holonomic basis we have $`{}_{}{}^{4}\overline{g}_{\overline{A}\overline{B}}^{}=ฯต\left(\begin{array}{cc}1& 0\\ 0& {}_{}{}^{3}g_{rs}^{}\end{array}\right)`$, $`{}_{}{}^{4}\overline{g}_{}^{\overline{A}\overline{B}}=ฯต\left(\begin{array}{cc}1& 0\\ 0& {}_{}{}^{3}g_{}^{rs}\end{array}\right)`$, $`\sqrt{{}_{}{}^{4}\overline{g}}=\sqrt{\gamma }/\sqrt{ฯต{}_{}{}^{4}\overline{g}_{}^{\overline{o}\overline{o}}}`$, $`\sqrt{{}_{}{}^{4}\overline{g}_{}^{1}}=\sqrt{{}_{}{}^{3}\overline{g}_{}^{1}}/\sqrt{ฯต{}_{}{}^{4}\overline{g}_{\overline{o}\overline{o}}^{}}=\sqrt{ฯต{}_{}{}^{4}\overline{g}_{\overline{o}\overline{o}}^{}}/\sqrt{{}_{}{}^{4}\overline{g}}=\sqrt{{}_{}{}^{4}\overline{g}_{}^{1}}`$\]. We get at $`z`$
$$n=\frac{J^o}{U^o\sqrt{{}_{}{}^{4}\overline{g}}}=\eta _{123}det_k\stackrel{~}{\alpha }^i\frac{\sqrt{ฯต{}_{}{}^{4}\overline{g}_{\overline{o}\overline{o}}^{}}}{\sqrt{{}_{}{}^{4}\overline{g}}}=\eta _{123}det_k\stackrel{~}{\alpha }^i\sqrt{det|{}_{}{}^{4}\overline{g}_{}^{rs}|}=\eta _{123}\sqrt{detG^{ij}}.$$
(546)
Therefore, we have $`n=\eta _{123}\sqrt{detG^{ij}}`$. Moreover, the material space of elastic materials, which have not only volume rigidity but also shape rigidity, is equipped with a Riemannian (symmetric and definite positive) 3-metric $`\gamma _{ij}^{(M)}(\alpha ^i)`$, the โmaterial metricโ, which is frozen in the material and it is not a dynamical object of the theory. It describes the โwould beโ local rest-frame space distance between neighbouring โmoleculesโ, measured in the locally relaxed state of the material. To measure the components $`\gamma _{ij}^{(M)}(\alpha ^i)`$ we have to relax the material at different points $`\alpha ^i(\tau ,\stackrel{}{\sigma })`$ separately, since global relaxation of the material may not be possible \[the material space may not be isometric with any 3-dimensional subspace of $`M^4`$, as in classical non-linear elastomechanics, when the material exhibits internal stresses frozen in it\]. The components $`\gamma _{ij}^{(M)}=\gamma _{ij}^{(M)}(\alpha ^i(\tau ,\stackrel{}{\sigma }))`$ are given functions, which describe axiomatically the properties of the material (the theory is fully invariant with respect to reparametrizations of the material space).
For ideal fluids $`\gamma _{ij}^{(M)}=\delta _{ij}`$; this also holds for non-pre-stressed materials without โinternalโ or โfrozenโ stresses.
Now the material space volume element $`\eta `$ has
$$\eta _{123}(\alpha ^i)=\sqrt{det\gamma _{ij}^{(M)}(\alpha ^i)},sothatn=\eta _{123}\sqrt{detG^{ij}}=\sqrt{det\gamma _{ij}^{(M)}}\sqrt{detG^{ij}},$$
(547)
and it cannot be put $`=1`$ like for perfect fluids.
The pull-back of the material metric $`\gamma _{ij}^{(M)}`$ to $`M^4`$ is
$$\gamma _{\stackrel{ห}{A}\stackrel{ห}{B}}^{(M)}=\gamma _{ij}^{(M)}_{\stackrel{ห}{A}}\alpha ^i_{\stackrel{ห}{B}}\alpha ^jsatisfying\gamma _{\stackrel{ห}{A}\stackrel{ห}{B}}^{(M)}U^{\stackrel{ห}{B}}=0.$$
(548)
The next step is to define a measure of the difference between the induced 3-metric $`G_{ij}(\alpha ^i)`$ and the constitutive metric $`\gamma _{ij}^{(M)}(\alpha ^i)`$, to be taken as a measure of the deformation of the material and as a definition of a โrelativistic strain tensorโ, locally vanishing when there is a local relaxation of the material. Some existing proposal for such a tensor in $`M^4`$ are:
$$i)S^{(1)}_{\stackrel{ห}{A}\stackrel{ห}{B}}=\frac{1}{2}({}_{}{}^{4}g_{\stackrel{ห}{A}\stackrel{ห}{B}}^{}ฯตU_{\stackrel{ห}{A}}U_{\stackrel{ห}{B}}\gamma _{\stackrel{ห}{A}\stackrel{ห}{B}}^{(M)})$$
(549)
which vanishes at relax and satisfies $`S_{\stackrel{ห}{A}\stackrel{ห}{B}}^{(1)}U^{\stackrel{ห}{B}}=0`$ \[but it must satisfy the involved matrix inequality $`2detS^{(1)}det({}_{}{}^{4}gฯตUU)`$\].
$$ii)S^{(2)\stackrel{ห}{A}}_{\stackrel{ห}{B}}=\frac{1}{2}(K^{\stackrel{ห}{A}}{}_{\stackrel{ห}{B}}{}^{}\delta _{\stackrel{ห}{B}}^{\stackrel{ห}{A}}),S^{(2)\stackrel{ห}{A}}_{\stackrel{ห}{B}}U^{\stackrel{ห}{B}}=0,$$
(550)
with $`K^{\stackrel{ห}{A}}{}_{\stackrel{ห}{B}}{}^{}={}_{}{}^{4}g_{}^{\stackrel{ห}{A}\stackrel{ห}{C}}(\gamma _{\stackrel{ห}{C}\stackrel{ห}{B}}^{(M)}ฯตU_{\stackrel{ห}{C}}U_{\stackrel{ห}{B}})`$, $`K^{\stackrel{ห}{A}}{}_{\stackrel{ห}{B}}{}^{}U_{}^{\stackrel{ห}{B}}=U^{\stackrel{ห}{A}}`$ \[the 4-velocity field is an eigenvector of the $`K`$-matrix\].
$$iii)S^{(3)}=\frac{1}{2}lnK,$$
(551)
with the same $`K`$-matrix as in ii).
However a simpler proposal is to define a โrelativistic strain tensorโ in the material space
$$S_i{}_{}{}^{j}=\gamma _{ik}^{(M)}G^{kj},$$
(552)
with locally $`S_i{}_{}{}^{j}=\delta _i^j`$ when there is local relaxation of the material (in this case physical spacelike distances between material points near a point $`z^\mu (\tau ,\stackrel{}{\sigma })`$ agree with their material distances). Since $`n=\eta _{123}\sqrt{detG^{ij}}=\sqrt{det\gamma _{ij}^{(M)}}\sqrt{detG^{ij}}`$, we have $`n=\sqrt{detS_i^j}`$ for the local rest-frame matter number density.
We can now define the local rest-frame energy per unit volume of the material $`n(\tau ,\stackrel{}{\sigma })e(\tau ,\stackrel{}{\sigma })`$, where $`e`$ denotes the molar local rest-frame energy (moles = number of particles)
$$e(\tau ,\stackrel{}{\sigma })=m+u_I(\tau ,\stackrel{}{\sigma }).$$
(553)
Here $`m`$ is the molar local rest mass, $`u_I`$ is the amount of internal energy (per mole of the material) of the elastic deformations, accumulated in an infinitesimal portion during the deformation from the locally relaxed state to the actual state of strain.
For isotropic media $`u_I`$ may depend on the deformation only via the invariants of the strain tensor.
Let us notice that for an anisotropic material (like a crystal) the energy $`u_I`$ may depend upon the orientation of the deformation with respect to a specific axis, reflecting the microscopic composition of the material: this information may be encoded in a vector field $`E^i(\tau ,\stackrel{}{\sigma })`$ in the material space and one may assume $`u_I=u_I(G_{ij}^1E^iE^j)`$.
The function $`e=e[\alpha ^i,G^{ij},\gamma _{ij}^{(M)},\mathrm{}]`$ describes the dependence of the energy of the material upon its state of strain and plays the role of an โequation of stateโ or โconstitutive equationโ of the material.
In the weak strain approximation of an isotropic elastic continuum (Hooke approximation) the function $`u_I`$ depends only on the linear ($`h=S_i^i`$) and quadratic ($`q=S_i{}_{}{}^{j}S_{}^{j}_i`$) invariants of the strain tensor and coincides with the standard formula of linear elasticity \[$`V=\frac{1}{n}=\frac{1}{\sqrt{detS_i^j}}`$ is the specific volume\],
$$u_I=\lambda (V)h^2+2\mu (V)q+O(cubicinvariants),$$
(554)
where $`\lambda `$ and $`\mu `$ are the Lamรฉ coefficients.
The action principle for this description of relativistic materials is
$$S[{}_{}{}^{4}g,\alpha ^i,\alpha ^i]=๐\tau d^3\sigma L(\tau ,\stackrel{}{\sigma })=๐\tau d^3\sigma (N\sqrt{\gamma })(\tau ,\stackrel{}{\sigma })n(\tau ,\stackrel{}{\sigma })e(\tau ,\stackrel{}{\sigma }).$$
(555)
It is shown in Refs. that the canonical stress-energy- momentum tensor $`T_{\stackrel{ห}{B}}^{\stackrel{ห}{A}}=p_i^{\stackrel{ห}{A}}_{\stackrel{ห}{B}}\alpha ^i\delta _{\stackrel{ห}{B}}^{\stackrel{ห}{A}}L`$, where $`p_i^{\stackrel{ห}{A}}=\frac{L}{_{\stackrel{ห}{A}}\alpha ^i}`$ is the relativistic Piola-Kirchhoff momentum density, and coincides with the symmetric energy-momentum tensor $`T_{\stackrel{ห}{A}\stackrel{ห}{B}}=2\frac{L}{{}_{}{}^{4}g_{}^{\stackrel{ห}{A}\stackrel{ห}{B}}}`$, which satisfies $`T^{\stackrel{ห}{A}\stackrel{ห}{B}}{}_{;\stackrel{ห}{B}}{}^{}=0`$ due to the Euler-Lagrange equations. This energy-momentum tensor may be written in the following form
$$T_{\stackrel{ห}{A}\stackrel{ห}{B}}=N\sqrt{\gamma }n[eU_{\stackrel{ห}{A}\stackrel{ห}{B}}+Z_{\stackrel{ห}{A}\stackrel{ห}{B}}],$$
(556)
where $`Z_{\stackrel{ห}{A}\stackrel{ห}{B}}=Z_{ij}_{\stackrel{ห}{A}}\alpha ^i_{\stackrel{ห}{B}}\alpha ^j`$ is the pull-back from the material space to $`M^4`$ of the โresponse tensorโ of the material
$$Z_{ij}=2\frac{e}{G^{ij}},sothatde(G)=\frac{1}{2}Z_{ij}dG^{ij}.$$
(557)
The part $`\tau _{\stackrel{ห}{A}\stackrel{ห}{B}}=nZ_{\stackrel{ห}{A}\stackrel{ห}{B}}`$, $`\tau _{\stackrel{ห}{A}\stackrel{ห}{B}}U^{\stackrel{ห}{B}}=0`$, may be called the relativistic โstress or Cauchyโ tensor and contains the โstress-strain relationโ through the dependence of $`Z_{ij}`$ on $`S_i^j`$ implied by the consitutive equation $`e=e[\alpha ,\gamma ^{(M)},G,..]`$ of the material.
For an isotropic elastic material we get
$$Z_{ij}=V\left[pG_{ij}^1+B\gamma _{ij}^{(M)}+CG_{ij}\right],$$
(558)
where $`p=\frac{e}{V}`$, $`B=\frac{2}{V}\frac{e}{h}`$, $`C=\frac{2}{V}\frac{e}{q}`$ and we get
$$de(V,h,q)=pdV+\frac{1}{2}VBdh+\frac{1}{2}VCdq.$$
(559)
The response parameters describe the reaction of the material to the strain: $`p`$ is the โisotropic stressโ, while $`B`$ and $`C`$ give the anisotropic response as in non-relativistic elesticity \[perfect fluids have $`e=e(V)`$, $`B=C=0`$, $`Z_{ij}=VpG_{ij}^1`$ and $`de(V)=pdV`$ is the Pascal law\].
See Ref. for a different description of relativistic Hooke law in linear elasticity: there is a 4-dimensional deformation tensor $`S_{\mu \nu }=\frac{1}{2}({}_{}{}^{4}_{\mu }^{}\xi _\nu +{}_{}{}^{4}_{\nu }^{}\xi _\mu )`$ and the constitutive equations of the material are given in the form $`T^{\mu \nu }=C^{(\mu \nu )(\alpha \beta )}S_{\alpha \beta }`$.
In Ref. the theory is also extended to the thermodynamics of isentropic flows (no heat conductivity). The function $`e`$ is considered also as a function of entropy $`S=S(\alpha )`$ and $`de=\frac{1}{2}Z_{ij}dG^{ij}`$ is generalized to
$$de=\frac{1}{2}Z_{ij}dG^{ij}SdT.$$
(560)
Then $`e`$ is replaced with the Helmholtz free energy $`f=eTS`$ so to obtain
$$df=\frac{1}{2}Z_{ij}dG^{ij}SdT,$$
(561)
\[for perfect fluids we get $`de(V,S)=pdV+TdS`$, $`df(V,T)=pdVSdT`$\]. This suggests to consider the temperature $`T`$ as a strain and the entropy $`S`$ as the corresponding stress and to introduce an extra scalar field $`\alpha ^\tau (\tau ,\stackrel{}{\sigma })`$ so that $`T=const.U^{\stackrel{ห}{A}}_{\stackrel{ห}{A}}\alpha ^\tau `$. The potential $`\alpha ^\tau (\tau ,\stackrel{}{\sigma })`$ has the microscopic interpretation as the retardation of the proper time of the molecules with respect to the physical time calculated over averaged spacetime trajectories of the idealized continuum material.
In this case the action principle becomes $`S=๐\tau d^3\sigma \left[N\sqrt{\gamma }nf(G,T)\right](\tau ,\stackrel{}{\sigma })`$ and one gets the conserved energy-momentum tensor $`T_{\stackrel{ห}{A}\stackrel{ห}{B}}=N\sqrt{\gamma }n[(f+TS)U_{\stackrel{ห}{A}}U_{\stackrel{ห}{B}}+Z_{\stackrel{ห}{A}\stackrel{ห}{B}}]`$.
In all these cases one can develop the rest-frame instant form just in the same way as it was done in Section II and III for perfect fluids, even if it is not possible to obtain a closed form of the invariant mass.
## VIII Conclusions.
In this paper we have studied the Hamiltonian description in Minkowski spacetime associated with an action principle for perfect fluids with an equation of state of the form $`\rho =\rho (n,s)`$ given in Ref., in which the fluid is descrbed only in terms of Lagrangian coordinates.
This action principle can be reformulated on arbitrary spacelike hypersurfaces embedded in Minkowski spacetime (covariant 3+1 splitting of Minkowski spacetime) along the lines of Refs.. At the Hamiltonian leve the canonical Hamiltonian vanishes and the theory is governed by four first class constraints $`^\mu (\tau ,\stackrel{}{\sigma })0`$ implying the independence of the description from the choice of the 3+1 splitting of Minkowski spacetime.
These constraints can be obtained in closed form only for the โdustโ and for the โphoton gasโ. For other types of perfect fluids one needs numerical calculations. After the inclusion of the coupling to the gravitational field one could begin to think to formulate Hamiltonian numerical gravity with only physical degrees of freedom and hyperbolic Hamilton equations for them \[like the form (175) of the relativistic Euler equations for the dust\].
After the canonical reduction to 3+1 splittings whose leaves are spacelike hyperplanes, we consider all the configurations of the perfect fluid whose conserved 4-momentum is timelike. For each of these configurations we can select the special foliation of Minkowski spacetime with spacelike hyperplanes orthogonal to the 4-momentum of the configuration,
This gives rise to the โWigner-covariant rest-frame instant form of dynamicsโ for the perfect fluids. After a discussion of the โexternalโ and โinternalโ centers of mass and realizations of the Poincarรฉ algebra, rest-frame Dixonโs Cartesian multipoles of the perfect fluid are studied.
It is also shown that the formulation of non-dissipative elastic materials of Ref., based on the use of Lagrangian coordinates, allows to get the rest-frame instant form for these materials too.
Finally it is shown how to make the coupling to the gravitational field by giving the ADM action for the perfect fluid in tetrad gravity. Now it becomes possible to study the canonical reduction of tetrad gravity with the perfect fluids as matter along the lines of Refs..
## A Relativistic Perfect Fluids.
As in Ref. let us consider a perfect fluid in a curved spacetime $`M^4`$ with unit 4-velocity vector field $`U^\mu (z)`$, Lagrangian coordinates $`\stackrel{~}{\alpha }^i(z)`$, particle number density $`n(z)`$, energy density $`\rho (z)`$, entropy per particle $`s(z)`$, pressure $`p(z)`$, temperature $`T(z)`$. Let $`J^\mu (z)=\sqrt{{}_{}{}^{4}g(z)}n(z)U^\mu (z)`$ the densitized particle number flux vector field, so that we have $`n=\sqrt{ฯต{}_{}{}^{4}g_{\mu \nu }^{}J^\mu J^\nu }/\sqrt{{}_{}{}^{4}g}`$. Other local thermodynamical variables are the chemical potential or specific enthalpy (the energy per particle required to inject a small amount of fluid into a fluid sample, keeping the sample volume and the entropy per particle $`s`$ constant)
$$\mu =\frac{1}{n}(\rho +p),$$
(A1)
the physical free energy (the injection energy at a constant number density n and constant total entropy)
$$a=\frac{\rho }{n}Ts,$$
(A2)
and the chemical free energy (the injection energy at constant volume and constant total entropy)
$$f=\frac{1}{n}(\rho +p)Ts=\mu Ts.$$
(A3)
Since the local expression of the first law of thermodynamics is
$$d\rho =\mu dn+nTds,ordp=nd\mu nTds,ord(na)=fdnnsdT,$$
(A4)
an equation of state for a perfect fluid may be given in one of the following forms
$$\rho =\rho (n,s),orp=p(\mu ,s),ora=a(n,T).$$
(A5)
By definition, the stress-energy-momentum tensor for a perfect fluid is
$$T^{\mu \nu }=ฯต\rho U^\mu U^\nu +p({}_{}{}^{4}g_{}^{\mu \nu }ฯตU^\mu U^\nu )=ฯต(\rho +p)U^\mu U^\nu +p{}_{}{}^{4}g_{}^{\mu \nu },$$
(A6)
and its equations of motion are
$$T^{\mu \nu }{}_{;\nu }{}^{}=0,(nU^\mu )_{;\mu }=\frac{1}{\sqrt{{}_{}{}^{4}g}}_\mu J^\mu =0.$$
(A7)
As shown in Ref. an action functional for a perfect fluid depending upon $`J^\mu (z)`$, $`{}_{}{}^{4}g_{\mu \nu }^{}(z)`$, $`s(z)`$ and $`\stackrel{~}{\alpha }^i(z)`$ requires the introduction of the following Lagrange multipliers to implement all the required properties: i) $`\theta (z)`$: it is a scalar field named โthermasyโ; it is interpreted as a potential for the fluid temperature $`T=\frac{1}{n}\frac{\rho }{s}|_n`$. In the Lagrangian it is interpreted as a Lagrange multiplier for implementing the โentropy exchange constraintโ $`(sJ^\mu )_{,\mu }=0`$. ii) $`\phi (z)`$: it is a scalar field; it is interpreted as a potential for the chemical free energy $`f`$. In the Lagrangian it is interpreted as a Lagrange multipliers for the โparticle number conservation constraintโ $`J^\mu {}_{,\mu }{}^{}=0`$. iii) $`\beta _i(z)`$: they are three scalar fields; in the Lagrangian they are interpreted as Lagrange multipliers for the โconstraintโ $`\stackrel{~}{\alpha }^i{}_{,\mu }{}^{}J_{}^{\mu }=0`$ that restricts the fluid 4-velocity vector to be directed along the flow lines $`\stackrel{~}{\alpha }^i=const.`$
Given an arbitrary equation of state of the type $`\rho =\rho (n,s)`$, the action functional is
$`S[{}_{}{}^{4}g_{\mu \nu }^{},J^\mu ,s,\stackrel{~}{\alpha },\phi ,\theta ,\beta _i]`$ $`=`$ $`{\displaystyle }d^4z\{\sqrt{{}_{}{}^{4}g}\rho ({\displaystyle \frac{|J|}{\sqrt{{}_{}{}^{4}g}}},s)+`$ (A8)
$`+`$ $`J^\mu [_\mu \phi +s_\mu \theta +\beta _i_\mu \stackrel{~}{\alpha }^i]\}.`$ (A9)
By varying the 4-metric we get the standard stress-energy-momentum tensor
$$T^{\mu \nu }=\frac{2}{\sqrt{{}_{}{}^{4}g}}\frac{\delta S}{\delta {}_{}{}^{4}g_{\mu \nu }^{}}=ฯต\rho U^\mu U^\nu +p({}_{}{}^{4}g_{}^{\mu \nu }ฯตU^\mu U^\nu )=ฯต(\rho +p)U^\mu U^\nu +p{}_{}{}^{4}g_{}^{\mu \nu },$$
(A10)
where the pressure is given by
$$p=n\frac{\rho }{n}|_s\rho .$$
(A11)
The Euler-Lagrange equations for the fluid motion are
$`{\displaystyle \frac{\delta S}{\delta J^\mu }}`$ $`=`$ $`\mu U_\mu +_\mu \phi +s_\mu \theta +\beta _i_\mu \stackrel{~}{\alpha }^i=0,`$ (A12)
$`{\displaystyle \frac{\delta S}{\delta \phi }}`$ $`=`$ $`_\mu J^\mu =0,`$ (A13)
$`{\displaystyle \frac{\delta S}{\delta \theta }}`$ $`=`$ $`_\mu (sJ^\mu )=0,`$ (A14)
$`{\displaystyle \frac{\delta S}{\delta s}}`$ $`=`$ $`\sqrt{{}_{}{}^{4}g}{\displaystyle \frac{\rho }{s}}+J^\mu _\mu \theta =0,`$ (A15)
$`{\displaystyle \frac{\delta S}{\delta \stackrel{~}{\alpha }^i}}`$ $`=`$ $`_\mu (\beta _iJ^\mu )=0,`$ (A16)
$`{\displaystyle \frac{\delta S}{\delta \beta _i}}`$ $`=`$ $`J^\mu _\mu \stackrel{~}{\alpha }^i=0.`$ (A17)
The second equation is the particle number conservation, the third one the entropy exchange constraint and the last one restricts the fluid 4-velocity vector to be directed along the flow lines $`\stackrel{~}{\alpha }^i=const.`$. The first equation gives the Clebsch or velocity-potential representation of the 4-velocity $`U_\mu `$ (the scalar fields in this representation are called Clebsch or velocity potentials). The fifth equations imply the constancy of the $`\beta _i`$โs along the fluid flow lines, so that these Lagrange multipliers can be expressed as a function of the Lagrangian coordinates. The fourth equation, after a comparison with the first law of thermodynamics, leads to the identification $`T=U^\mu _\mu \theta =\frac{1}{n}\frac{\rho }{s}|_n`$ for the fluid temperature.
Moreover, one can show that the Euler-Lagrange equations imply the conservation of the stress-energy-momentum tensor $`T^{\mu \nu }{}_{;\nu }{}^{}=0`$. This equations can be split in the projection along the fluid flow lines and in the one orthogonal to them: i) The projection along the fluid flow lines plus the particle number conservation give $`U_\mu T^{\mu \nu }{}_{;\nu }{}^{}=\frac{\rho }{s}U^\mu _\mu s=0`$, which is verified due to the entropy exchange constraint. Therefore, the fluid flow is locally adiabatic, that is the entropy per particle along the fluid flow lines is conserved. ii) The projection orthogonal to the fluid flow lines gives the Euler equations, relating the fluid acceleration to the gradient of pressure
$$({}_{}{}^{4}g_{\mu \nu }^{}ฯตU_\mu U_\nu )T^{\nu \alpha }{}_{;\alpha }{}^{}=ฯต(\rho +p)U_{\mu ;\nu }U^\nu (\delta _\mu ^\nu ฯตU_\mu U^\nu )_\nu p.$$
(A18)
By using $`p=n\frac{\rho }{n}|_s\rho `$, it is shown in Ref. that these equations can be rewritten as
$$2(\mu U_{[\mu })_{;\nu ]}U^\nu =ฯต(\delta _\mu ^\nu U_\mu U^\nu )\frac{1}{n}\frac{\rho }{s}|_n_\nu s.$$
(A19)
The use of the entropy exchange constraint allows the rewrite the equations in the form
$$2V_{[\mu ;\nu ]}U^\nu =T_\mu s,$$
(A20)
where $`V_\mu =\mu U_\mu `$ is the Taub current (important for the description of circulation and vorticity), which can be identified with the 4-momentum per particle of a small amount of fluid to be injected in a larger sample of fluid without changing the total fluid volume or the entropy per particle. Now from the Euler-Lagrange we get
$$2V_{[\mu ;\nu ]}U^\nu =2(_{[\mu }\phi +s_{[\mu }\theta +\beta _i_{[\mu }\stackrel{~}{\alpha }^i)_{;\nu ]}U^\nu =(s_{[\mu }\theta )_{;\nu ]}U^\nu =T_\mu s,$$
(A21)
and this result implies the validity of the Euler equations.
In the non-relativistic limit $`(nU^\mu )_{;\mu }=0`$, $`T^{\mu \nu }{}_{;\nu }{}^{}=0`$ become the particle number (or mass) conservation law, the entropy conservation law and the Euler-Newton equations. See Refs. for the post-Newtonian approximation.
We refer to Ref. for the complete discussion. The previous action has the advantage on other actions that the canonical momenta conjugate to $`\phi `$ and $`\theta `$ are the particle number density and entropy density seen by Eulerian observers at rest in space. The action evaluated on the solutions of the equations of motion is $`d^4z\sqrt{{}_{}{}^{4}g(z)}p(z)`$.
In Ref. there is a study of a special class of global Noether symmetries of this action associated with arbitrary functions $`F(\stackrel{~}{\alpha },\beta _i,s)`$. It is shown that for each $`F`$ there is a conservation equation $`_\mu (FJ^\mu )=0`$ and a Noether charge $`Q[F]=_\mathrm{\Sigma }d^3\sigma \sqrt{\gamma }n(ฯตl_\mu U^\mu )F(\stackrel{~}{\alpha },\beta _i,s)`$ \[$`\mathrm{\Sigma }`$ is a spacelike hypersurface with future pointing unit normal $`l_\mu `$ and with a 3-metric with determinant $`\sqrt{\gamma }`$\]. For $`F=1`$ inside a volume V in $`\mathrm{\Sigma }`$ we get the conservation of particle number within a flow tube defined by the bundle of flow lines contained in the volume V. The factor $`ฯตl_\mu U^\mu `$ is the relativistic โgamma factorโ characterizing a boost from the Lagrangian observers with 4-velocity $`U^\mu `$ to the Eulerian observers with 4-velocity $`l^\mu `$; thus $`n(ฯตl_\mu U^\mu )`$ is the particle number density as seen from the Eulerian observers. These symmetries describe the changes of Lagrangian coordinates $`\stackrel{~}{\alpha }^i`$ and the fact that both the Lagrange multipliers $`\phi `$ and $`\theta `$ are constant along each flow line (so that it is possible to transform any solution to the fluid equations of motion into a solution with $`\phi =\theta =0`$ on any given spacelike hypersurface).
However, the Hamiltonian formulation associated with this action is not trivial, because the many redundant variables present in it give rise to many first and second class constraints. In particular we get: 1) second class constraints: A) $`\pi _{J^\tau }0`$, $`J^\tau \pi _\phi 0`$; B) $`\pi _s0`$, $`sJ^\tau \pi _\theta 0`$; C) $`\pi _{\beta _s}0`$, $`\beta _sJ^\tau \pi _{\alpha ^s}0`$. 2) first class constraints: $`\pi _{J^r}0`$, so that the $`J^r`$โs are gauge variables. Therefore the physical variables are the five pairs: $`\phi `$, $`\pi _\phi `$; $`\theta `$, $`\pi _\theta `$; $`\stackrel{~}{\alpha }^i`$, $`\pi _{\alpha ^r}`$ and one could study the associated canonical reduction.
In Ref. \[see its rich bibliography for the references\] there is a systematic study of the action principles associated to the three types of equations of state present in the literature, first by using the Clebsch potentials and the associated Lagrange multipliers, then only in terms of the Lagrangian coordinates by inserting the solution of some of the Euler-Lagrange equations in the original action and eventually by adding surface terms.
1) Equation of state $`\rho =\rho (n,s)`$. One has the action
$$S[n,U^\mu ,\phi ,\theta ,s,\stackrel{~}{\alpha }^r,\beta _r;{}_{}{}^{4}g_{\mu \nu }^{}]=d^4x\sqrt{{}_{}{}^{4}g}\left[\rho (n,s)nU^\mu (_\mu \phi \theta _\mu s+\beta _r_\mu \stackrel{~}{\alpha }^r)\right]$$
(A22)
If one knows $`s=s(\stackrel{~}{\alpha }^r)`$ and $`J^\mu =J^\mu (\stackrel{~}{\alpha }^r)=\sqrt{{}_{}{}^{4}g}ฯต^{\mu \nu \rho \sigma }_\nu \stackrel{~}{\alpha }^1_\rho \stackrel{~}{\alpha }^2_\sigma \stackrel{~}{\alpha }^3\eta _{123}(\stackrel{~}{\alpha }^r)`$, one can define $`\stackrel{~}{S}=Sd^4x_\mu [(\phi +s\theta )J^\mu ]`$, and one can show that it has the form
$$\stackrel{~}{S}=\stackrel{~}{S}[\stackrel{~}{\alpha }^r]=d^4x\sqrt{{}_{}{}^{4}g}\rho (\frac{|J|}{\sqrt{{}_{}{}^{4}g}},s).$$
(A23)
2) Equation of state: $`p=p(\mu ,s)`$ \[$`V^\mu =\mu U^\mu `$ Taub vector\]
$`S_{(p)}`$ $`=`$ $`s_{(p)}[V^\mu ,\phi ,\theta ,s,\stackrel{~}{\alpha }^r,\beta _r;{}_{}{}^{4}g_{\mu \nu }^{}]=`$ (A24)
$`=`$ $`{\displaystyle d^4x\sqrt{{}_{}{}^{4}g}\left[p(\mu ,s)\frac{p}{\mu }\left(|V|\frac{V^\mu }{|V|}(_\mu \phi +s_\mu \theta +\beta _r_\mu \stackrel{~}{\alpha }^r)\right)\right]}`$ (A25)
or by using one of its EL equations $`V_\mu \stackrel{}{=}(_\mu \phi +s_\mu \theta +\beta _r_\mu \stackrel{~}{\alpha }^r)`$ to eliminate $`V^\mu `$ one gets Schutzโs action \[$`\mu `$ determined by $`\mu ^2=V^\mu V_\mu `$\]
$$\stackrel{~}{S}_{(p)}[\phi ,\theta ,s,\stackrel{~}{\alpha }^r,\beta _r;{}_{}{}^{4}g_{\mu \nu }^{}]=d^4x\sqrt{{}_{}{}^{4}g}p(\mu ,s)$$
(A26)
3) Equation of state $`a=a(n,T)`$. The action is
$$S_{(a)}[J^\mu ,\phi ,\theta ,\stackrel{~}{\alpha }^r,\beta _r;{}_{}{}^{4}g_{\mu \nu }^{}]=d^4x\left[|J|a(\frac{|J|}{\sqrt{{}_{}{}^{4}g}},_\mu \theta J^\mu )J^\mu (_\mu \phi +\beta _r_\mu \stackrel{~}{\alpha }^r)\right]$$
(A27)
or $`\stackrel{~}{S}_{(a)}[\phi ,\theta ,s,\stackrel{~}{\alpha }^r,\beta _r]=S_{(a)}d^4x\left[|J|a(\frac{|J|}{\sqrt{{}_{}{}^{4}g}},\frac{J^\mu }{|J|}_\mu \theta )\right]`$.
At the end of Ref. there is the action for โisentropicโ fluids and for their particular case of a โdustโ (used in Ref. as a reference fluid in canonical gravity).
The isentropic fluids have equation of state $`a(n,T)=\frac{\rho (n)}{n}sT`$ with $`s=const.`$ (constant value of the entropy per particle). By introducing $`\phi ^{^{}}=\phi +s\theta `$, the action can be written in the form
$$S_{(isentrpic)}[J^\mu ,\phi ^{^{}},\stackrel{~}{\alpha }^r,\beta _r,{}_{}{}^{4}g_{}^{\mu \nu }]=d^4x\left[\sqrt{{}_{}{}^{4}g}\rho (\frac{|J|}{\sqrt{{}_{}{}^{4}g}})+J^\mu (_\mu \phi ^{^{}}+\beta _r_\mu \stackrel{~}{\alpha }^r)\right]$$
(A28)
or
$$\stackrel{~}{S}_{(isentropic)}[\stackrel{~}{\alpha }^r;{}_{}{}^{4}g_{\mu \nu }^{}]=d^4x\sqrt{{}_{}{}^{4}g}\rho (\frac{|J|}{\sqrt{{}_{}{}^{4}g}})$$
(A29)
The dust has equation of state $`\rho (n)=\mu n`$, namely $`a(n,T)=\mu sT`$ so that we get zero pressure $`p=n\frac{\rho }{n}\rho =0`$. Again with $`\phi ^{^{}}=\phi +s\theta `$ the action becomes
$$S_{(dust)}J^\mu ,\phi ^{^{}},\stackrel{~}{\alpha }^r,\beta _r,{}_{}{}^{4}g_{}^{\mu \nu }]=d^4x[\mu |J|+J^\mu (_\mu \phi ^{^{}}+\beta _r_\mu \stackrel{~}{\alpha }^r)]$$
(A30)
or with $`U_\mu =\frac{1}{\mu }(_\mu \phi ^{^{}}+\beta _r_\mu \stackrel{~}{\alpha }^r)`$ \[In Ref.: $`M=\mu n`$ rest mass (energy) density and $`T=\phi ^{^{}}/\mu `$, $`W_r=\beta _r`$, $`Z^r=\stackrel{~}{\alpha }^r`$; $`U_\mu =_\mu T+W_r_\mu Z^r`$\]
$$S_{(dust)}^{^{}}[T,Z^r,M,W_r;{}_{}{}^{4}g_{\mu \nu }^{}]=\frac{1}{2}d^4x\sqrt{{}_{}{}^{4}g}(\mu n)\left(U_\mu {}_{}{}^{4}g_{}^{\mu \nu }U_\nu ฯต\right),$$
(A31)
or
$$\stackrel{~}{S}_{(dust)}[\stackrel{~}{\alpha }^r;{}_{}{}^{4}g_{\mu \nu }^{}]=d^4x\mu |J|$$
(A32)
In Ref. there is a study of the action (A31) since the dust is used as a reference fluid in general relativity. At the Hamiltonian level one gets: i) 3 pairs of second class constraints \[$`\pi _\stackrel{}{W}^r(\tau ,\stackrel{}{\sigma })0`$, $`\pi _{\stackrel{}{Z}r}(\tau ,\stackrel{}{\sigma })W_r(\tau ,\stackrel{}{\sigma })\pi _T(\tau ,\stackrel{}{\sigma })0`$\], which allow the elimination of $`W_r(\tau ,\stackrel{}{\sigma })`$ and $`\pi _\stackrel{}{W}^r(\tau ,\stackrel{}{\sigma })`$; ii) a pair of second class constraints \[ $`\pi _M(\tau ,\stackrel{}{\sigma })0`$ plus the secondary $`M(\tau ,\stackrel{}{\sigma })\frac{\pi _T^2}{\sqrt{\gamma }\sqrt{\pi _T^2+{}_{}{}^{3}g_{}^{rs}(\pi _T_rT+\pi _{\stackrel{}{Z}u}_rZ^u)(\pi _T_sT+\pi _{\stackrel{}{Z}v}_sZ^v)}}(\tau ,\stackrel{}{\sigma })0`$\], which allow the elimination of $`M,\pi _M`$.
## B Covariant Relativistic Thermodynamics of Equilibrium and Non-Equilibrium.
In this Appendix we shall collect some results on relativistic fluids which are well known but scattered in the specialized literature. We shall use essentially Ref., which has to be consulted for the relevant bibliography. See also Ref..
Firstly we remind some notions of covariant thermodynamics of equilibrium.
Let us remember that given the stress-energy-momentum tensor of a continuous medium $`T^{\mu \nu }`$, the densities of energy and momentum are $`T^{oo}`$ and $`c^1T^{ro}`$ respectively \[so that $`dP^\mu =c^1\eta T^{\mu \nu }d\mathrm{\Sigma }_\nu `$ is the 4-momentum that crosses the 3-area element $`d\mathrm{\Sigma }_\nu `$ in the sense of its normal ($`\eta =1`$ if the normal is spacelike, $`\eta =+1`$ if it is timelike)\]; instead, $`cT^{or}`$ is the energy flux in the positive $`r`$ direction, while $`T^{rs}`$ is the $`r`$ component of the stress in the plane perpendicular to the $`s`$ direction (a pressure, if it is positive). A local observer with timelike 4-velocity $`u^\mu `$ ($`u^2=ฯตc^2`$) will measure energy density $`c^2T^{\mu \nu }u_\mu u_\nu `$ and energy flux $`ฯตT^{\mu \nu }u_\mu n_\nu `$ along the direction of a unit vector $`n^\mu `$ in his rest frame.
For a fluid at thermal equilibrium with $`T^{\mu \nu }=\rho U^\mu U^\nu ฯต\frac{p}{c^2}({}_{}{}^{4}g_{}^{\mu \nu }ฯตU^\mu U^\nu )`$ \[$`U^\mu `$ is the hydrodynamical 4-velocity of the fluid\] with particle number density $`n`$, specific volume $`V=\frac{1}{n}`$ and entropy per particle $`s=\frac{k_BS}{n}`$ ($`k_B`$ is Boltzmannโs constant) in its rest frame, the energy density is
$$\rho c^2=n(mc^2+e),$$
(B1)
where $`e`$ is the mean internal (thermal plus chemical) energy per particle and $`m`$ is particleโs rest mass.
From a non-relativistic point of view, by writing the equation of state in the form $`s=s(e,V)`$ the temperature and the pressure emerge as partial derivatives from the first law of thermodynamics in the form (Gibbs equation)
$$ds(e,V)=\frac{1}{T}(de+pdV).$$
(B2)
If $`\mu _{clas}=e+pVTs`$ is the non-relativistic chemical potential per particle, its relativistic version is
$$\mu ^{^{}}=mc^2+\mu _{class}=\mu Ts,$$
(B3)
\[$`\mu =\frac{\rho c^2+p}{n}`$ is the specific enthalpy, also called chemical potential as in Appendix A\] and we get
$`\mu ^{^{}}n`$ $`=`$ $`\rho c^2+pnTs=\rho c^2+pk_BTS,`$ (B4)
$`k_BTdS`$ $`=`$ $`d(\rho c^2)\mu ^{^{}}dn=d(\rho c^2)(\mu Ts)dn,or`$ (B5)
$`d(\rho c^2)`$ $`=`$ $`\mu ^{^{}}dn+Td(ns)=\mu dn+nTds.`$ (B6)
By introducing the โthermal potentialโ $`\alpha =\frac{\mu ^{^{}}}{k_BT}=\frac{\mu Ts}{k_BT}`$ and the inverse temperature $`\beta =\frac{c^2}{k_BT}`$, these two equations take the form
$`S`$ $`=`$ $`{\displaystyle \frac{ns}{k_B}}=\beta (\rho +{\displaystyle \frac{p}{c^2}})\alpha n,`$ (B7)
$`dS`$ $`=`$ $`\beta d\rho \alpha dn.`$ (B8)
Let us remark that in Refs. one uses different notations, some of which are given in the following equation (in Ref. $`\rho `$ is denoted $`e`$ and $`\rho ^{^{}}`$ is denoted $`r`$)
$$\rho +\frac{p}{c^2}=n(mc^2+e)+\frac{p}{c^2}=\rho ^{^{}}h=\rho ^{^{}}(c^2+e^{^{}}+\frac{p}{c^2\rho ^{^{}}})=\rho ^{^{}}(c^2+h^{^{}}),$$
(B9)
where $`\rho ^{^{}}=nm`$ is the rest-mass density \[$`r_{}=\sqrt{{}_{}{}^{4}g}\rho ^{^{}}`$ is called the coordinate rest-mass density\] and $`e^{^{}}=e/m`$ is the specific internal energy \[so that $`\rho ^{^{}}h`$ is the โeffective inertial mass of the fluid; in the post-Newtonian approximation of Ref. it is shown that $`\sigma =c^2(T^{oo}+_sT^{ss})+O(c^4)=c^2\sqrt{{}_{}{}^{4}g}(T_o^o+T_s^s)+O(c^4)`$ has the interpretation of equality of the โpassiveโ and the โactiveโ gravitational mass\]. For the specific enthalpy or chemical potential we get \[$`\mu /m=h=c^2+h^{^{}}`$ is called enthalpy\]
$$\mu =\frac{1}{n}(\rho +\frac{p}{c^2})=\frac{m}{\rho ^{^{}}}(\rho +\frac{p}{c^2})=mh=m(c^2+h^{^{}}).$$
(B10)
See Ref. for a richer table of conversion of notations.
Relativistically, we must consider, besides the stress-energy-momentum tensor $`T^{\mu \nu }`$ and the associated 4-momentum $`P^\mu =_Vd^3\mathrm{\Sigma }_\nu T^{\mu \nu }`$, a particle flux density $`n^\mu `$ (one $`n_a^\mu `$ for each constituent $`a`$ of the system) and the entropy flux density $`s^\mu `$. At thermal equilibrium all these a priori unrelated 4-vectors must all be parallel to the hydrodynamical 4-velocity
$$n^\mu =nU^\mu ,s^\mu =sU^\mu ,P^\mu =PU^\mu ,$$
(B11)
Analogously, we have $`V^\mu =VU^\mu `$ ($`V=1/n`$ is the specific volume), $`\beta ^\mu =\beta U^\mu =\frac{c^2}{k_BT}U^\mu `$ \[a related 4-vector is the equilibrium parameter 4-vector $`i^\mu =\mu ^{^{}}\beta ^\mu `$\].
Since $`ฯตU_\mu T^{\mu \nu }=\rho U^\nu `$, we get the final manifestly covariant form of the previous two equations (now the hydrodynamical 4-velocity is considered as an extra thermodynamical variable)
$`S^\mu `$ $`=`$ $`SU^\mu ={\displaystyle \frac{ns}{k_B}}U^\mu ={\displaystyle \frac{p}{c^2}}\beta ^\mu \alpha n^\mu ฯต\beta _\nu T^{\nu \mu },`$ (B12)
$`dS^\mu `$ $`=`$ $`\alpha dn^\mu ฯต\beta _\nu dT^{\nu \mu }.`$ (B13)
Global thermal equilibrium imposes
$$_\mu \alpha =_\mu \beta _\nu +_\nu \beta _\mu =0.$$
(B14)
As a consequence we get
$$d(\frac{p}{c^2}\beta ^\mu )=n^\mu d\alpha +ฯตT^{\nu \mu }d\beta _\nu ,$$
(B15)
namely the basic variables $`n^\mu `$, $`T^{\nu \mu }`$ and $`S^\mu `$ can all be generated from partial derivatives of the โfugacityโ 4-vector (or โthermodynamical potentialโ)
$`\varphi ^\mu (\alpha ,\beta _\lambda )`$ $`=`$ $`{\displaystyle \frac{p}{c^2}}\beta ^\mu ,`$ (B16)
$`n^\mu `$ $`=`$ $`{\displaystyle \frac{\varphi ^\mu }{\alpha }},T_{(mat)}^{\nu \mu }={\displaystyle \frac{\varphi ^\mu }{\beta _\nu }},S^\mu =\varphi ^\mu \alpha n^\mu \beta _\nu T_{(mat)}^{\nu \mu },`$ (B17)
once the equation of state is known. Here $`T_{(mat)}^{\nu \mu }`$ is the canonical or material (in general non symmetric) stress tensor, ensuring that reversible flows of field energy are not accompanied by an entropy flux.
This final form remains valid (at least to first order in deviations) for โstates that deviate from equilibriumโ, when the 4-vectors $`S^\mu `$, $`n^\mu `$,โฆ are no more parallel; the extra information in this equation is precisely the standard linear relation between entropy flux and heat flux. The second law of thermodynamics for relativistic systems is $`_\mu S^\mu 0`$, which becomes a strict equality in equilibrium.
The fugacity 4-vector $`\varphi ^\mu `$ is evaluated by using the covariant relativistic statistical theory for thermal equilibrium starting from a grand canonical ensemble with density matrix $`\widehat{\rho }`$ by maximizing the entropy $`S=Tr(\widehat{\rho }ln\widehat{\rho })`$ subject to the constraints $`Tr\widehat{\rho }=1`$, $`Tr(\widehat{\rho }\widehat{n})=n`$, $`Tr(\widehat{\rho }\widehat{P}^\lambda )=P^\lambda `$: this gives (in the large volume limit)
$`\widehat{\rho }`$ $`=`$ $`Z^1e^{\alpha \widehat{n}+\beta _\mu \widehat{P}^\mu },`$ (B19)
$`with`$
$`lnZ`$ $`=`$ $`{\displaystyle _\mathrm{}\mathrm{\Sigma }}ฯต\varphi ^\mu ๐\mathrm{\Sigma }_\mu ,n={\displaystyle _\mathrm{}\mathrm{\Sigma }}ฯตn^\mu ๐\mathrm{\Sigma }_\mu ,P^\mu ={\displaystyle _\mathrm{}\mathrm{\Sigma }}ฯตT_{(mat)}^{\mu \nu }๐\mathrm{\Sigma }_\nu ,`$ (B20)
\[it is assumed that the members of the ensemble are small (macroscopic) subregions of one extended body in thermal equilibrium, whose worldtubes intersect an arbitrary spacelike hypersurface in small 3-areas $`\mathrm{}\mathrm{\Sigma }`$\]. Therefore, one has to find the grand canonical partition function
$$Z(V_\mu ,\beta _\mu ,i_\mu )=\underset{n}{}e^{i_\mu n^\mu }Q_n(V_\mu ,\beta _\mu ),whereQ_n(V_\mu ,\beta _\mu )=_{V_\mu }๐\sigma _n(q,p)e^{\beta _\mu P^\mu },$$
(B21)
is the canonical partition function for fixed volume $`V_\mu `$ and $`d\sigma _n`$ is the invariant microcanonical density of states. For an ideal Boltzmann gas of N free particles of mass $`m`$ \[see Section III for its equation of state\] it is
$$d\sigma _n(p,m)=\frac{1}{N!}\delta ^4(P\underset{i=1}{\overset{N}{}}p_i)\underset{i=1}{\overset{N}{}}2V_\mu p_i^\mu \theta (p_i^o)\delta (p_i^2ฯตm^2)d^4p_i.$$
(B22)
Following Ref. \[using a certain type of gauge fixings to the first class constraints $`p_i^2ฯตm^20`$\] in Ref. $`Q_n`$ was evaluated in the rest-frame instant form on the Wigner hyperplane (this method can be extended to a gas of molecules, which are N-body bound states):
$$Q_n=\frac{1}{N!}\left[\frac{Vm^2}{2\pi ^2\beta }K_2(m\beta )\right]^N.$$
(B23)
The same results may be obtained by starting from the covariant relativistic kinetic theory of gas \[see Ref.; in Ref. there is a short review\] whose particles interact only by collisions by using Syngeโs invariant distribution function $`N(q,p)`$ \[the number of particle worldlines with momenta in the range $`(p_\mu ,d\omega )`$ that cross a target 3-area $`d\mathrm{\Sigma }_\mu `$ in $`M^4`$ in the direction of its normal is given by $`dN=N(q,p)d\omega \eta v^\mu d\mathrm{\Sigma }_\mu `$ ($`=Nd^3qd^3p`$ for the 3-space $`q^o=const.`$); $`d\omega =d^3p/v^o\sqrt{{}_{}{}^{4}g}`$ is the invariant element of 3-area on the mass-shell\]. One arrives at a transport equation for $`N`$,$`\frac{dN}{d\tau }=_\mu (Nv^\mu )`$ \[$`v^\mu `$ is the particle velocity obtained from the Hamilton equation implied by the one-particle Hamiltonian $`H=\sqrt{ฯต{}_{}{}^{4}g_{}^{\mu \nu }(q)p_\mu p_\nu }=m`$ (it is the energy after the gauge fixing $`q^o\tau `$ to the first class contraint $`{}_{}{}^{4}g_{}^{\mu \nu }(q)p_\mu p_\nu ฯตm^20`$); $`_\mu `$ is the covariant gradient holding the 4-vector $`p_\mu `$ (not its components) fixed\] with a โcollision termโ $`C[N]`$ describing the collisions; for a dilute simple gas dominated by binary collisions one arrives at the Boltzmann equation \[for $`C[N]=0`$ one solution is the relativistic version of the Maxwell-Boltzmann distribution function, i.e. the classical Jรผttner-Synge one $`N=const.e^{\beta _\mu P^\mu }/4\pi m^2K_2(m\beta )`$ for the Boltzmann gas\]. The H-theorem \[$`_\mu S^\mu 0`$, where $`S^\mu (q)=[Nln(Nh^3)N]v^\mu ๐\omega `$ is the entropy flux\] and the results at thermal equilibrium emerge \[from the balance law $`_\mu (Nfv^\mu ๐\omega )=fC[N]๐\omega `$ ($`f`$ is an arbitrary tensorial function) one can deduce the conservation laws $`_\mu n^\mu =_\mu T^{\nu \mu }=0`$, where $`n^\mu =Nv^\mu ๐\omega `$, $`T_\nu {}_{}{}^{\mu }=Np_\nu v^\mu d\omega `$; the vanishing of entropy production at local thermal equilibrium gives $`N_{eq}(q,p)=h^3e^{\alpha (q)+\beta _\nu (q)P^\nu }`$ in the case of Boltzmann statistic and one gets ($`U^\mu =\beta ^\mu /\beta `$) $`n_{eq}^\mu =N_{eq}v^\mu ๐\omega =nU^\mu `$, $`T_{eq}^{\nu \mu }=\rho U^\nu U^\mu ฯตp({}_{}{}^{4}g_{}^{\nu \mu }ฯตU^\nu U^\mu )`$, $`S_{eq}^\mu =p\beta ^\mu \alpha n_{eq}^\mu \beta _\nu T_{eq}^{\mu \nu }`$; one obtains the equations for the Boltzmann ideal gas given in Section III\].
One can study the small deviations from thermal equilibrium \[$`N=N_{eq}(1+f)`$, where $`N_{eq}`$ is an arbitrary local equilibrium distribution\] with the linearized Boltzmann equation and then by using either the Chapman-Enskog ansatz of quasi-stationarity of small deviations (this ignores the gradients of $`f`$ and gives the standard Landau-Lifshitz and Eckart phenomenological laws; one gets Fourier equation for heat conduction and the Navier-Stokes equation for the bulk and shear stresses; however one has parabolic and not hyperbolic equations implying non-causal propagation) or with the Grad method in the 14-moment approximation. This method retains the gradients of $`f`$ \[there are 5 extra thermodynamical variables, which can be explicitly determined from 14 moments among the infinite set of moments $`Np^\mu p^\nu p^\rho \mathrm{}d^4p`$ of kinetic theory; no extra auxiliary state variables are introduced to specify a non-equilibrium state besides $`T^{\mu \nu }`$, $`n^\mu `$, $`S^\mu `$\] and gives phenomenological laws which are the kinetic equivalent of Mรผller extended thermodynamics and its various developments; now the equations are hyperbolic, there is no causality problem but there are problems with shock waves. See Ref. for the bibliography and for a review of the non-equilibrium phenomenological laws (see also Ref. ) of Eckart, Landau-Lifshitz, of the various formulations of extended thermodynamics, of non-local thermodynamics.
While in Ref. it is said that the difference between causal hyperbolic theories and acausual parabolic one is unobservable, in Ref. \[see also Ref.\] there is a discussion of the cases in which hyperbolic theories are relevant. See also the numerical codes of Refs..
In phenomenological theories the starting point are the equations $`_\mu T^{\mu \nu }=_\mu n^\mu =0`$, $`_\mu S^\mu 0`$. There is the problem of how to define a 4-velocity and a rest-frame for a given non-equilibrium state. Another problem is how to specify a non-equilibrium state completely at the macroscopic level: a priori one could need an infinite number of auxiliary quantities (vanishing at equilibrium) and an equation of state depending on them. The basic postulate of extended thermodynamics is the absence of such variables.
Regarding the rest frame problem there are two main solutions in the literature connected with the relativistic description of โheat flowโ:
i) Eckart theory. One considers a local observer in a simple fluid who is at rest with respect to the average motion of the particles: its 4-velocity $`U_{(eck)}^\mu `$ is parallel by definition to the particle flux $`n^\mu `$, namely
$$n^\mu =n_{(eck)}U_{(eck)}^\mu .$$
(B24)
This local observer sees โheat flowโ as a flux of energy in his rest frame:
$`ฯตU_{(eck)\mu }T^{\mu \nu }`$ $`=`$ $`\rho _{(eck)}U_{(eck)}^\mu +q_{(eck)}^\mu ,`$ (B26)
$`sothatweget`$
$`T^{\mu \nu }`$ $`=`$ $`\rho _{(eck)}U_{(eck)}^\mu U_{(eck)}^\nu +q_{(eck)}^\mu U_{(eck)}^\nu +U_{(eck)}^\mu q_{(eck)}^\nu +P_{(eck)}^{\mu \nu },`$ (B27)
$`P_{(eck)}^{\mu \nu }`$ $`=`$ $`P_{(eck)}^{\nu \mu }=ฯต(p+\pi _{(eck)})({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu +\pi _{(eck)}^{\mu \nu },`$ (B31)
$`P_{(eck)}^{\mu \nu }U_{(eck)\nu }=q_{(eck)\mu }U_{(eck)}^\mu =0,\pi _{(eck)\mu \nu }({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )=0,`$
where $`p`$ is the thermodynamic pressure, $`\pi _{(eck)}`$ the bulk viscosity and $`\pi _{(eck)}^{\mu \nu }`$ the shear stress.
This description has the particle conservation law $`_\mu n^\mu =0`$.
ii) Landau-Lifshitz theory. One considers a different observer (drifting slowly in the direction of heat flow with a 3-velocity $`\stackrel{}{v}_D=\stackrel{}{q}/nmc^2`$) whose 4-velocity $`U_{(ll)}^\mu `$ is by definition such to give a vanishing โheat flowโ, i.e. there is no net energy flux in his rest frame: $`U_{(ll)}^\mu T_\mu ^\nu n_\nu =0`$ for all vectors $`n_\mu `$ orthogonal to $`U_{(ll)}^\mu `$. This implies that $`U_{(ll)}^\mu `$ is the timelike eigenvector of $`T^{\mu \nu }`$, $`T^{\mu \nu }U_{(ll)\nu }=ฯต\rho _{(ll)}U_{(ll)}^\mu `$, which is unique if $`T^{\mu \nu }`$ satisfies a positive energy condition. Now we get
$`T^{\mu \nu }`$ $`=`$ $`\rho _{(ll)}U_{(ll)}^\mu U_{(ll)}^\nu +P_{(ll)}^{\mu \nu },`$ (B32)
$`P_{(ll)}^{\mu \nu }`$ $`=`$ $`P_{(ll)}^{\nu \mu }=ฯต(p+\pi _{(ll)})({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu +\pi _{(ll)}^{\mu \nu },`$ (B36)
$`P_{(ll)}^{\mu \nu }U_{(ll)\nu }=0,\pi _{(ll)\mu \nu }({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu )=0,`$
$`n^\mu `$ $`=`$ $`n_{(ll)}U_{(ll)}^\mu +j_{(ll)}^\mu ,j_{(ll)\mu }U_{(ll)}^\mu =0(\stackrel{}{j}=n\stackrel{}{v}_D=\stackrel{}{q}/mc^2).`$ (B38)
This observer in his rest frame does not see a heat flow but a particle drift. This description has the simplest form of the energy-momentum tensor.
One has $`n_{(eck)}=n_{(ll)}ch\phi `$, $`\rho _{(eck)}=\rho _{(ll)}ch^2\phi +p_{(ll)}sh^2\phi =\pi ^{\mu \nu }j_\mu j_\nu /n_{(eck)}^2`$, with $`ch\phi =U_{(ll)}^\mu U_{(eck)\mu }`$ \[the difference is a Lorentz factor $`\sqrt{1\stackrel{}{v}_D^2/c^2}`$, so that there are insignificant differences for many practical purposes if deviations from equilibrium are small\]. The angle $`\phi j/nv_D/cq/nmc^2`$ is a dimensionless measure of the deviation from equilibrium \[$`n_{(ll)}n_{(eck)}`$ and $`\rho _{(ll)}\rho _{(eck)}`$ are of order $`\phi ^2`$\].
One can decompose $`T^{\mu \nu }`$, $`n^\mu `$, in terms of any 4-velocity $`U^\mu `$ that falls within a cone of angle $`\phi `$ containing $`U_{(eck)}^\mu `$ and $`U_{(ll)}^\mu `$; each choice $`U^\mu `$ gives a particle density $`n(U)=ฯตu_\mu n^\mu `$ and energy density $`\rho (U)=U_\mu U_\nu T^{\mu \nu }`$ which are independent of $`U^\mu `$ if one neglects terms of order $`\phi ^2`$. Therefore, one has: i) if $`S_{eq}(\rho (U),n(U))`$ is the equilibrium entropy density, then $`S(U)=ฯตU_\mu S^\mu =S_{eq}+O(\phi ^2)`$; ii) if $`p(U)=\frac{\rho /n}{1/n}|_{S/n}`$ is the (reversible) thermodynamical pressure defined as work done in an isentropic expansion (off equilibrium this definition allows to sepate it from the bulk stress $`\pi (U)`$ in the stress-energy-momentum tensor) and $`p_{eq}`$ is the pressure at equilibrium, then $`p(U)=P_{eq}(\rho (U),n(U))+O(\phi ^2)`$.
By postulating that the covariant Gibbs relation remains valid for arbitrary infinitesimal displacements $`(\delta n^\mu ,\delta T^{\mu \nu },..)`$ from an equilibrium state, one gets a covariant off-equilibrium thermodynamics based on the equation
$`S^\mu `$ $`=`$ $`p(\alpha ,\beta )\beta ^\mu \alpha n^\mu \beta _\nu T^{\mu \nu }Q^\mu (\delta n^\nu ,\delta T^{\nu \rho },..),`$ (B39)
$`_\mu S^\mu `$ $`=`$ $`\delta n^\mu _\mu \alpha \delta T^{\mu \nu }_\nu \beta _\mu _\mu Q^\mu 0,`$ (B40)
with $`Q^\mu `$ of second order in the displacements and $`\alpha `$, $`\beta _\mu `$ arbitrary. At equilibrium one recovers $`S_{eq}^\mu =p\beta ^\mu \alpha n_{eq}^\mu \beta _\nu T_{eq}^{\mu \nu }`$, $`_\mu S_{eq}^\mu =0`$ \[with $`U^\mu =\beta ^\mu /\beta `$ and (for viscous heat-conducting fluids, but not for superfluids) $`_\mu \alpha =_\mu \beta _\nu +_\nu \beta _\mu =0`$\].
If we choose $`\beta ^\mu =U^\mu /k_BT`$ parallel to $`n^\mu `$ of the given off-equilibrium state, we are in the โEckart frameโ, $`U^\mu =U_{(eck)}^\mu `$, and we get
$`S`$ $`=`$ $`ฯตU_{(eck)\mu }S^\mu =S_{eq}+ฯตU_{(eck)\mu }Q^\mu ,`$ (B41)
$`\sigma _{(eck)}^\mu `$ $`=`$ $`({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )S_\nu =\beta q_{(eck)}^\mu ({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )Q_\nu ,`$ (B43)
$`q_{(eck)}^\mu `$ $`=`$ $`({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )T_\nu ^\rho U_{(eck)\rho },`$ (B45)
so that to linear order we get the standard relation between entropy flux $`\stackrel{}{\sigma }_{(eck)}`$ and heat flux $`\stackrel{}{q}_{(eck)}`$
$$\stackrel{}{\sigma }_{(eck)}=\frac{\stackrel{}{q}_{(eck)}}{k_BT}+(possible\mathrm{\hspace{0.17em}\hspace{0.17em}2}ndorderterm).$$
(B46)
If we choose $`U^\mu =U_{(ll)}^\mu `$, the timelike eigenvector of $`T^{\mu \nu }`$, so that $`U_{(ll)\mu }T_\nu ^\mu ({}_{}{}^{4}g_{\rho }^{\nu }ฯตU_{(ll)}^\nu U_{(ll)\rho })=0`$, we are in the โLandau-Lifshitz frameโ and we get
$`\sigma _{(ll)}^\mu `$ $`=`$ $`({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu )S_\nu =\alpha j_{(ll)}^\mu ({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu )Q_\nu ,`$ (B47)
$`j_{(ll)}^\mu `$ $`=`$ $`({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu )n_\nu ,`$ (B49)
so that at linear order we get the standard relation between entropy flux $`\stackrel{}{\sigma }_{(ll)}`$ and diffusive flux $`\stackrel{}{j}_{(ll)}`$
$$\stackrel{}{\sigma }_{(ll)}=\frac{\mu }{k_BT}\stackrel{}{j}_{(ll)}+(possible\mathrm{\hspace{0.17em}\hspace{0.17em}2}ndorderterm).$$
(B50)
In the Landau-Lifschitz frame heat flow and diffusion are englobed in the diffusive flux $`\stackrel{}{j}_{(ll)}`$ relative to the mean mass-energy flow. The entropy inequality becomes (each term is of second order in the deviations from local equilibrium)
$$0_\mu S^\mu =\delta n^\mu _\mu \alpha \delta T^{\mu \nu }_\nu \beta _\mu _\mu Q^\mu ,$$
(B51)
with the fitting conditions $`\delta n^\mu U_\mu =\delta T^{\mu \nu }U_\mu U_\nu =0`$, which contain all information about the viscous stresses, heat flow and diffusion in the off-equilibrium state (they are dependent on the arbitrary choice of the 4-velocity $`U^\mu `$). Once a detailed form of $`Q^\mu `$ is specified, linear relations between irreversible fluxes $`\delta T^{\mu \nu }`$, $`\delta n^\mu `$ and gradients $`_{(\mu }\beta _{\nu )}`$, $`_\mu \alpha `$ follow. A) $`Q^\mu =0`$ (like in the non-relativistic case). The spatial entropy flux $`\stackrel{}{\sigma }`$ is only a strictly linear function of heat flux $`\stackrel{}{q}`$ and diffusion flux $`\stackrel{}{j}`$. In this case the off-equilibrium entropy density $`S=ฯตU_\mu S^\mu `$ is given by the equilibrium equation of state $`S=S_{eq}(\rho ,n)`$. We have $`0_\mu S^\mu =\delta n^\mu _\mu \alpha \delta T^{\mu \nu }_\nu \beta _\mu `$ with fitting conditions $`\delta n^\mu U_\mu =\delta T^{\mu \nu }U_\mu U_\nu =0`$ and with $`U^\mu `$ still arbitrary at first order. A1) Landau-Lifschitz frame and theory. $`U^\mu =U_{(ll)}^\mu `$ is the timelike eigenvector of $`T^{\mu \nu }`$. This and the fitting conditions imply $`\delta T^{\mu \nu }U_{(ll)\nu }=0`$. The shear and bulk stresses $`\pi _{(ll)}^{\mu \nu }`$, $`\pi _{(ll)}`$ are identified by the decomposition
$$\delta T^{\mu \nu }=\pi _{(ll)}^{\mu \nu }+\pi _{(ll)}({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu ),\pi _{(ll)}^{\mu \nu }U_{(ll)\nu }=\pi _{(ll)}^\mu {}_{\mu }{}^{}=0.$$
(B52)
The inequality $`_\mu S^\mu 0`$ becomes
$`j_{(ll)}^\mu _\mu \alpha \beta \pi _{(ll)}^{\mu \nu }<_\nu \beta _\mu >\beta \pi _{(ll)}_\mu U_{(ll)}^\mu 0,j_{(ll)}^\mu =\delta n^\mu ,`$ (B53)
$`<X_{\mu \nu }>`$ $`=`$ $`[({}_{}{}^{4}g_{\mu }^{\alpha }ฯตU_{(ll)}^\alpha U_{(ll)\mu })({}_{}{}^{4}g_{\nu }^{\beta }ฯตU_{(ll)}^\beta U_{(ll)\nu })`$ (B55)
$``$ $`{\displaystyle \frac{1}{3}}({}_{}{}^{4}g_{\mu \nu }^{}ฯตU_{(ll)\mu }U_{(ll)\nu })({}_{}{}^{4}g_{}^{\alpha \beta }ฯตU_{(ll)}^\alpha U_{(ll)}^\beta )]X_{\alpha \beta },`$ (B56)
\[the $`<..>`$ operation extracts the purely spatial, trace-free part of any tensor\]. If the equilibrium state is isotropic (Curieโs principle) and if we assume that $`(j_{(ll)}^\mu ,\pi _{(ll)}^{\mu \nu },\pi _{(ll)})`$ are โlinear and purely localโ functions of the gradiants, $`_\mu S^\mu 0`$ implies
$$j_{(ll)}^\mu =\kappa ({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(ll)}^\mu U_{(ll)}^\nu )_\mu \alpha ,\kappa >0,$$
(B57)
\[it is a mixture of Fourierโs law of heat conduction and of Fickโs law of diffusion, stemming from the relativistic mass-energy equivalence\], and the standard Navier-Stokes equations ($`\zeta _S`$, $`\zeta _V`$ are shear and bulk viscosities)
$$\pi _{(ll)\mu \nu }=2\zeta _S<_\nu \beta _\mu >,\pi _{(ll)}=\frac{1}{3}\zeta _V_\mu U_{(ll)}^\mu .$$
(B58)
A2) Eckart frame and theory. $`U_{(eck)}^\mu `$ parellel to $`n^\mu `$. Now we have the fitting condition $`\delta n^\mu =0`$. The heat flux appears in the decomposition of $`\delta T^{\mu \nu }`$ \[$`a_{(eck)\mu }=U_{(eck)}^\nu _\nu U_{(eck)\mu }`$ is the 4-acceleration\]
$$\delta T^{\mu \nu }=q_{(eck)}^\mu U_{(eck)}^\nu +U_{(eck)}^\mu q_{(eck)}^\nu +\pi _{(eck)}^{\mu \nu }+\pi _{(eck)}({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu ).$$
(B59)
The inequality $`_\mu S^\mu 0`$ becomes
$$q_{(eck)}^\mu (_\mu \alpha \beta a_{(eck)\mu })\beta (\pi _{(eck)}^{\mu \nu }_\nu U_{(eck)\mu }+\pi _{(eck)}_\mu U_{(eck)}^\mu )0.$$
(B60)
With the simplest assumption of linearity and locality, we obtain Fourierโs law of heat conduction \[it is not strictly equivalent to the Landau-Lifshitz one, because they differ by spatial gradients of the viscous stresses and the time-derivative of the heat flux\]
$$q_{(eck)}^\mu =\kappa ({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )(_\nu T+T_\tau U_{(eck)\nu }),$$
(B61)
\[the term depending on the acceleration is sometimes referred to as an effect of the โinertia of heatโ\], and the same form of the Navier-Stokes equations for $`\pi _{(eck)}^{\mu \nu }`$, $`\pi _{(eck)}`$ (they are not strictly equivalent to the Landau-Lifshitz ones, because they differ by gradients of the drift $`\stackrel{}{v}_D=\stackrel{}{q}/nmc^2`$). For a simple fluid Fourierโs law and Navier-Stokes equations (9 equations) and the conservation laws $`_\mu T^{\mu \nu }=_\mu n^\mu =0`$ (5 equations) determine the 14 variables $`T^{\mu \nu }`$, $`n^\mu `$ from suitable initial data. However, these equations are of mixed parabolic-hyperbolic- elliptic type and, as said, one gets acausality and instability. Kinetic theory gives
$$Q^\mu =\frac{1}{2}N_{eq}f^2p^\mu ๐\omega 0,$$
(B62)
for a gas up to second order in the deviation $`(NN_{eq})=N_{eq}f`$ \[$`Q^\mu =0`$ requires small gradients and quasi-stationary processes\]. Two alternative classes of phenomenological theories are B) Linear non-local thermodynamics (NLT). This theory gives a rheomorphic rather than causal description of the phenomenological laws: transport coefficients at an event $`x`$ are taken to depend, not on the entire causal past of $`x`$, but only on the past history of a โcomoving local fluid elementโ. It is a linear theory restricted to small deviations from equilibrium, which can be derived from the linearized Boltzmann equation by projector-operator techniques (and probably inherits its causality properties). Instead of writing $`(\delta n_\mu (x),\delta T_{\mu \nu }(x))=\sigma (U,T)(_\mu \alpha (x),_{(\mu }\beta _{\nu )}(x))`$, this local phenomenological law is generalized to $`(\delta n_\mu (\stackrel{}{x},x^o),\delta T_{\mu \nu }(\stackrel{}{x},x^o))=_{\mathrm{}}^{\mathrm{}}๐x^o^{}\sigma (x^ox^o^{})(_\mu \alpha (\stackrel{}{x}x^o^{}),_{(\mu }\beta _{\nu )}(\stackrel{}{x}x^o^{}))`$. C) Local non-linear extended thermodynamics (ET). It is more relevant for relativistic astrophysics, where correlation and memory effects are not of primary interest and, instead, one needs a tractable and consistent transport theory coextensive at the macroscopic level with Boltzmannโs equation. It is assumed that the second order term $`Q^\mu (\delta n^\nu ,\delta T^{\nu \rho },..)`$ does not depend on auxiliary variables vanishing at equilibrium: this ansatz is the phenomenological equivalent of Gradโs 14-moment approximation in kinetic theory. These theories are called โsecond-order theoriesโ and many of them are analyzed in Ref.; when the dissipative fluxes are subject to a conservation equation, these theories are called of causal โdivergence typeโ like the ones of Refs.. Another type of theory (extended irreversible thermodynamics; in general these theories are not of divergence type) was developed in Refs.: in it there are transport equations for the dissipative fluxes rather than conservation laws.
For small deviations one retains only the quadratic terms in the Taylor expansion of $`Q^\mu `$ (leading to โlinearโ phenomenological laws): this implies 5 new undetermined coefficients
$$Q^\mu =\frac{1}{2}U^\mu [\beta _o\pi ^2+\beta _1q^\mu q_\mu +\beta _2\pi ^{\mu \nu }\pi _{\mu \nu }]\alpha _o\pi q^\mu \alpha _1\pi ^{\mu \nu }q_\nu ,$$
(B63)
with $`\beta _i>0`$ from $`ฯตU_\mu Q^\mu >0`$ (the $`\beta _i`$โs are โrelaxation timesโ). A first-order change of rest frame produces a second-order change in $`Q^\mu `$ \[going from the Landau-Lifshitz frame to the Eckart one, one gets $`\alpha _{(eck)i}\alpha _{(ll)i}=\beta _{(ll)1}\beta _{(eck)1}=[(\rho +p)T]^1`$, $`\beta _{(ll)0}=\beta _{(eck)o}`$, $`\beta _{(ll)2}=\beta _{(eck)2}`$, and the phenomenological laws are now invariant to first order\]. In the โEckart frameโ the phenomenological laws take the form
$`q_{(eck)}^\mu `$ $`=`$ $`\kappa T({}_{}{}^{4}g_{}^{\mu \nu }ฯตU_{(eck)}^\mu U_{(eck)}^\nu )[T^1_\nu T+a_{(eck)\nu }+\beta _{(eck)1}_\tau q_{(eck)\nu }`$ (B64)
$``$ $`\alpha _{(eck)o}_\nu \pi _{(eck)}\alpha _{(eck)1}_\rho \pi _{(eck)\nu }^\rho ],`$ (B65)
$`\pi _{(eck)\mu \nu }`$ $`=`$ $`2\zeta _S[<_\nu U_{(eck)\mu }>+\beta _{(eck)2}_\tau \pi _{(eck)\mu \nu }\alpha _{(eck)1}<_\nu q_{(eck)\mu }>],`$ (B67)
$`\pi _{(eck)}`$ $`=`$ $`{\displaystyle \frac{1}{3}}\zeta _V[_\mu U_{(eck)}^\mu +\beta _{(eck)o}_\tau \pi _{(eck)}\alpha _{(eck)o}_\mu q_{(eck)}^\mu ],`$ (B69)
which reduce to the equation of the standard Eckart theory if the 5 relaxation ($`\beta _i`$) and coupling ($`\alpha _i`$) coefficients are put equal to zero. See for instance Ref. for a complete treatment and also Ref. . For appropriate values of these coefficients these equations are hyperbolic and, therefore, causal and stable. The transport equations can be understood as evolution equations for the dissipative variables as they describe how these fluxes evolve from an initial arbitrary state to a final steady one \[the time parameter $`\tau `$ is usually interpreted as the relaxation time of the dissipative processes\]. In the case of a gas the new coefficients can be found explicitly \[see also Ref. for a recent approach to relativistic interacting gases starting from the Boltzmann equation\], and they are purely thermodynamical functions. Wave front speeds are finite and comparable with the speed of sound. A problem with these theories is that they do not admit a regular shock structure (like the Navier-Stokes equations) once the speed of the shock front exceeds the highest characteristic velocity (a โsubshockโ will form within a shock layer for speeds exceeding the wave-front velocities of thermo-viscous effects). The situation slowly ameliorates if more moments are taken into account . In the approach reviewed in Ref. the extra indeterminacy associated to the new 5 coefficients is eliminated (at the price of high non-linearity) by annexing to the usual conservation and entropy laws a new phenomenological assumption (in this way one obtains a causal divergence type theory):
$$_\rho A^{\mu \nu \rho }=I^{\mu \nu },$$
(B70)
in which $`A^{\rho \mu \nu }`$ and $`I^{\mu \nu }`$ are symmetric tensors with the following traces
$$A^{\mu \nu }{}_{\nu }{}^{}=n^\mu ,I^\mu {}_{\mu }{}^{}=0.$$
(B71)
These conditions are modelled on kinetic theory, in which $`A^{\rho \mu \nu }`$ represents the third moment of the distribution function in momentum space, and $`I^{\mu \nu }`$ the second moment of the collision term in Boltzmannโs equation. The previous equations are central in the determination of the distribution function in Gradโs 14-moment approximation. The phenomenological theory is completed by the postulate that the state variables $`S^\mu `$, $`A^{\rho \mu \nu }`$, $`I^{\mu \nu }`$ are invariant functions of $`T^{\mu \nu }`$, $`n^\mu `$ only. The theory is an almost exact phenomenological counterpart of the Grad approximation. See Ref. for the beginning (only non viscous heta conducting materials are treated) of a derivation of extended thermodynamics from a variational principle.
Everything may be rephrased in terms of the Lagrangian coordinates of the fluid used in this paper. What is lacking in the non-dissipative case of heat conduction is the functional form of the off-equilibrium equation of state reducing to $`\rho =\rho (n,s)`$ at thermal equilibrium. In the dissipative case the system is open and $`T^{\mu \nu }`$, $`P^\mu `$, $`n^\mu `$ are not conserved.
See Ref. for attempts to define a classical theory of dissipation in the Hamiltonian framework and Ref. about Hamiltonian molecular dynamics for the addition of an extra degree of freedom to an N-body system to transform it into an open system (with the choice of a suitable potential for the extra variable the equilibrium distribution function of the N-body subsystem is exactly the canonical ensemble).
However, the most constructive procedure is to get (starting from an action principle) the Hamiltonian form of the energy-momentum of a closed system, like it has been done in Ref. for a system of N charged scalar particles, in which the mutual action-at-a-distance interaction is the complete Darwin potential extracted from the Lienard-Wiechert solution in the radiation gauge (the interactions are momentum- and, therefore, velocity-dependent). In this case one can define an open (in general dissipative) subsystem by considering a cluster of $`n<N`$ particles and assigning to it a non-conserved energy-momentum tensor built with all the terms of the original energy-momentum tensor which depend on the canonical variables of the $`n`$ particles (the other $`Nn`$ particles are considered as external fields).
## C Notations on spacelike hypersurfaces.
Let us first review some preliminary results from Refs. needed in the description of physical systems on spacelike hypersurfaces.
Let $`\{\mathrm{\Sigma }_\tau \}`$ be a one-parameter family of spacelike hypersurfaces foliating Minkowski spacetime $`M^4`$ with 4-metric $`\eta _{\mu \nu }=ฯต(+)`$, $`ฯต=\pm `$ \[$`ฯต=+1`$ is the particle physics convention; $`ฯต=1`$ the general relativity one\] and giving a 3+1 decomposition of it. At fixed $`\tau `$, let $`z^\mu (\tau ,\stackrel{}{\sigma })`$ be the coordinates of the points on $`\mathrm{\Sigma }_\tau `$ in $`M^4`$, $`\{\stackrel{}{\sigma }\}`$ a system of coordinates on $`\mathrm{\Sigma }_\tau `$. If $`\sigma ^{\stackrel{ห}{A}}=(\sigma ^\tau =\tau ;\stackrel{}{\sigma }=\{\sigma ^{\stackrel{ห}{r}}\})`$ \[the notation $`\stackrel{ห}{A}=(\tau ,\stackrel{ห}{r})`$ with $`\stackrel{ห}{r}=1,2,3`$ will be used; note that $`\stackrel{ห}{A}=\tau `$ and $`\stackrel{ห}{A}=\stackrel{ห}{r}=1,2,3`$ are Lorentz-scalar indices\] and $`_{\stackrel{ห}{A}}=/\sigma ^{\stackrel{ห}{A}}`$, one can define the vierbeins
$$z_{\stackrel{ห}{A}}^\mu (\tau ,\stackrel{}{\sigma })=_{\stackrel{ห}{A}}z^\mu (\tau ,\stackrel{}{\sigma }),_{\stackrel{ห}{B}}z_{\stackrel{ห}{A}}^\mu _{\stackrel{ห}{A}}z_{\stackrel{ห}{B}}^\mu =0,$$
(C1)
so that the metric on $`\mathrm{\Sigma }_\tau `$ is
$`g_{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })=z_{\stackrel{ห}{A}}^\mu (\tau ,\stackrel{}{\sigma })\eta _{\mu \nu }z_{\stackrel{ห}{B}}^\nu (\tau ,\stackrel{}{\sigma }),ฯตg_{\tau \tau }(\tau ,\stackrel{}{\sigma })>0,`$ (C2)
$`g(\tau ,\stackrel{}{\sigma })=detg_{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })=(detz_{\stackrel{ห}{A}}^\mu (\tau ,\stackrel{}{\sigma }))^2,`$ (C3)
$`\gamma (\tau ,\stackrel{}{\sigma })=detg_{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })=det{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}(\tau ,\stackrel{}{\sigma }),`$ (C4)
where $`g_{\stackrel{ห}{r}\stackrel{ห}{s}}=ฯต{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ with $`{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}`$ having positive signature $`(+++)`$.
If $`\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })=ฯต{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}`$ is the inverse of the 3-metric $`g_{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })`$ \[$`\gamma ^{\stackrel{ห}{r}\stackrel{ห}{u}}(\tau ,\stackrel{}{\sigma })g_{\stackrel{ห}{u}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })=\delta _{\stackrel{ห}{s}}^{\stackrel{ห}{r}}`$\], the inverse $`g^{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })`$ of $`g_{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })`$ \[$`g^{\stackrel{ห}{A}\stackrel{ห}{C}}(\tau ,\stackrel{}{\sigma })g_{\stackrel{ห}{c}\stackrel{ห}{b}}(\tau ,\stackrel{}{\sigma })=\delta _{\stackrel{ห}{B}}^{\stackrel{ห}{A}}`$\] is given by
$`g^{\tau \tau }(\tau ,\stackrel{}{\sigma })={\displaystyle \frac{\gamma (\tau ,\stackrel{}{\sigma })}{g(\tau ,\stackrel{}{\sigma })}},`$ (C7)
$`g^{\tau \stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })=[{\displaystyle \frac{\gamma }{g}}g_{\tau \stackrel{ห}{u}}\gamma ^{\stackrel{ห}{u}\stackrel{ห}{r}}](\tau ,\stackrel{}{\sigma })=ฯต[{\displaystyle \frac{\gamma }{g}}g_{\tau \stackrel{ห}{u}}{}_{}{}^{3}g_{}^{\stackrel{ห}{u}\stackrel{ห}{r}}](\tau ,\stackrel{}{\sigma }),`$
$`g^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })=\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })+[{\displaystyle \frac{\gamma }{g}}g_{\tau \stackrel{ห}{u}}g_{\tau \stackrel{ห}{v}}\gamma ^{\stackrel{ห}{u}\stackrel{ห}{r}}\gamma ^{\stackrel{ห}{v}\stackrel{ห}{s}}](\tau ,\stackrel{}{\sigma })=`$
$`=`$ $`ฯต{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })+[{\displaystyle \frac{\gamma }{g}}g_{\tau \stackrel{ห}{u}}g_{\tau \stackrel{ห}{v}}{}_{}{}^{3}g\stackrel{ห}{u}\stackrel{ห}{r}{}_{}{}^{3}g_{}^{\stackrel{ห}{v}\stackrel{ห}{s}}](\tau ,\stackrel{}{\sigma }),`$ (C8)
so that $`1=g^{\tau \stackrel{ห}{C}}(\tau ,\stackrel{}{\sigma })g_{\stackrel{ห}{C}\tau }(\tau ,\stackrel{}{\sigma })`$ is equivalent to
$$\frac{g(\tau ,\stackrel{}{\sigma })}{\gamma (\tau ,\stackrel{}{\sigma })}=g_{\tau \tau }(\tau ,\stackrel{}{\sigma })\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma })g_{\tau \stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })g_{\tau \stackrel{ห}{s}}(\tau ,\stackrel{}{\sigma }).$$
(C9)
We have
$$z_\tau ^\mu (\tau ,\stackrel{}{\sigma })=(\sqrt{\frac{g}{\gamma }}l^\mu +g_{\tau \stackrel{ห}{r}}\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}z_{\stackrel{ห}{s}}^\mu )(\tau ,\stackrel{}{\sigma }),$$
(C10)
and
$`\eta ^{\mu \nu }`$ $`=`$ $`z_{\stackrel{ห}{A}}^\mu (\tau ,\stackrel{}{\sigma })g^{\stackrel{ห}{A}\stackrel{ห}{B}}(\tau ,\stackrel{}{\sigma })z_{\stackrel{ห}{B}}^\nu (\tau ,\stackrel{}{\sigma })=`$ (C11)
$`=`$ $`(l^\mu l^\nu +z_{\stackrel{ห}{r}}^\mu \gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}z_{\stackrel{ห}{s}}^\nu )(\tau ,\stackrel{}{\sigma }),`$ (C12)
where
$`l^\mu (\tau ,\stackrel{}{\sigma })`$ $`=`$ $`({\displaystyle \frac{1}{\sqrt{\gamma }}}ฯต^\mu {}_{\alpha \beta \gamma }{}^{}z_{\stackrel{ห}{1}}^{\alpha }z_{\stackrel{ห}{2}}^\beta z_{\stackrel{ห}{3}}^\gamma )(\tau ,\stackrel{}{\sigma }),`$ (C14)
$`l^2(\tau ,\stackrel{}{\sigma })=1,l_\mu (\tau ,\stackrel{}{\sigma })z_{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma })=0,`$
is the unit (future pointing) normal to $`\mathrm{\Sigma }_\tau `$ at $`z^\mu (\tau ,\stackrel{}{\sigma })`$.
For the volume element in Minkowski spacetime we have
$`d^4z`$ $`=`$ $`z_\tau ^\mu (\tau ,\stackrel{}{\sigma })d\tau d^3\mathrm{\Sigma }_\mu =d\tau [z_\tau ^\mu (\tau ,\stackrel{}{\sigma })l_\mu (\tau ,\stackrel{}{\sigma })]\sqrt{\gamma (\tau ,\stackrel{}{\sigma })}d^3\sigma =`$ (C15)
$`=`$ $`\sqrt{g(\tau ,\stackrel{}{\sigma })}d\tau d^3\sigma .`$ (C16)
Let us remark that according to the geometrical approach of Ref.,one can use Eq.(C10) in the form $`z_\tau ^\mu (\tau ,\stackrel{}{\sigma })=N(\tau ,\stackrel{}{\sigma })l^\mu (\tau ,\stackrel{}{\sigma })+N^{\stackrel{ห}{r}}(\tau ,\stackrel{}{\sigma })z_{\stackrel{ห}{r}}^\mu (\tau ,\stackrel{}{\sigma })`$, where $`N=\sqrt{g/\gamma }=\sqrt{g_{\tau \tau }\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}g_{\tau \stackrel{ห}{r}}g_{\tau \stackrel{ห}{s}}}=\sqrt{g_{\tau \tau }+ฯต{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}g_{\tau \stackrel{ห}{r}}g_{\tau \stackrel{ห}{s}}}`$ and $`N^{\stackrel{ห}{r}}=g_{\tau \stackrel{ห}{s}}\gamma ^{\stackrel{ห}{s}\stackrel{ห}{r}}=ฯตg_{\tau \stackrel{ห}{s}}{}_{}{}^{3}g_{}^{\stackrel{ห}{s}\stackrel{ห}{r}}`$ are the standard lapse and shift functions $`N_{[z](flat)}`$, $`N_{[z](flat)}^{\stackrel{ห}{r}}`$ of the Introduction, so that $`g_{\tau \tau }=ฯตN^2+g_{\stackrel{ห}{r}\stackrel{ห}{s}}N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}}=ฯต[N^2{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}}]`$, $`g_{\tau \stackrel{ห}{r}}=g_{\stackrel{ห}{r}\stackrel{ห}{s}}N^{\stackrel{ห}{s}}=ฯต{}_{}{}^{3}g_{\stackrel{ห}{r}\stackrel{ห}{s}}^{}N^{\stackrel{ห}{s}}`$, $`g^{\tau \tau }=ฯตN^2`$, $`g^{\tau \stackrel{ห}{r}}=ฯตN^{\stackrel{ห}{r}}/N^2`$, $`g^{\stackrel{ห}{r}\stackrel{ห}{s}}=\gamma ^{\stackrel{ห}{r}\stackrel{ห}{s}}+ฯต\frac{N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}}}{N^2}=ฯต[{}_{}{}^{3}g_{}^{\stackrel{ห}{r}\stackrel{ห}{s}}\frac{N^{\stackrel{ห}{r}}N^{\stackrel{ห}{s}}}{N^2}]`$, $`\frac{}{z_\tau ^\mu }=l_\mu \frac{}{N}+z_{\stackrel{ห}{s}\mu }\gamma ^{\stackrel{ห}{s}\stackrel{ห}{r}}\frac{}{N^{\stackrel{ห}{r}}}=l_\mu \frac{}{N}ฯตz_{\stackrel{ห}{s}\mu }{}_{}{}^{3}g_{}^{\stackrel{ห}{s}\stackrel{ห}{r}}\frac{}{N^{\stackrel{ห}{r}}}`$, $`d^4z=N\sqrt{\gamma }d\tau d^3\sigma `$.
The rest frame form of a timelike fourvector $`p^\mu `$ is $`\stackrel{}{p}{}_{}{}^{\mu }=\eta \sqrt{ฯตp^2}(1;\stackrel{}{0})=\eta ^{\mu o}\eta \sqrt{ฯตp^2}`$, $`\stackrel{}{p}{}_{}{}^{2}=p^2`$, where $`\eta =signp^o`$. The standard Wigner boost transforming $`\stackrel{}{p}^\mu `$ into $`p^\mu `$ is
$`L^\mu {}_{\nu }{}^{}(p,\stackrel{}{p})`$ $`=`$ $`ฯต_\nu ^\mu (u(p))=`$ (C17)
$`=`$ $`\eta _\nu ^\mu +2{\displaystyle \frac{p^\mu \stackrel{}{p}_\nu }{ฯตp^2}}{\displaystyle \frac{(p^\mu +\stackrel{}{p}^\mu )(p_\nu +\stackrel{}{p}_\nu )}{p\stackrel{}{p}+ฯตp^2}}=`$ (C18)
$`=`$ $`\eta _\nu ^\mu +2u^\mu (p)u_\nu (\stackrel{}{p}){\displaystyle \frac{(u^\mu (p)+u^\mu (\stackrel{}{p}))(u_\nu (p)+u_\nu (\stackrel{}{p}))}{1+u^o(p)}},`$ (C19)
$`\nu =0`$ $`ฯต_o^\mu (u(p))=u^\mu (p)=p^\mu /\eta \sqrt{ฯตp^2},`$ (C21)
$`\nu =r`$ $`ฯต_r^\mu (u(p))=(u_r(p);\delta _r^i{\displaystyle \frac{u^i(p)u_r(p)}{1+u^o(p)}}).`$ (C22)
The inverse of $`L^\mu {}_{\nu }{}^{}(p,\stackrel{}{p})`$ is $`L^\mu {}_{\nu }{}^{}(\stackrel{}{p},p)`$, the standard boost to the rest frame, defined by
$$L^\mu {}_{\nu }{}^{}(\stackrel{}{p},p)=L_\nu {}_{}{}^{\mu }(p,\stackrel{}{p})=L^\mu {}_{\nu }{}^{}(p,\stackrel{}{p})|_{\stackrel{}{p}\stackrel{}{p}}.$$
(C23)
Therefore, we can define the following vierbeins \[the $`ฯต_r^\mu (u(p))`$โs are also called polarization vectors; the indices r, s will be used for A=1,2,3 and $`\overline{o}`$ for $`A=o`$\]
$`ฯต_A^\mu (u(p))=L^\mu {}_{A}{}^{}(p,\stackrel{}{p}),`$ (C24)
$`ฯต_\mu ^A(u(p))=L^A{}_{\mu }{}^{}(\stackrel{}{p},p)=\eta ^{AB}\eta _{\mu \nu }ฯต_B^\nu (u(p)),`$ (C25)
(C26)
$`ฯต_\mu ^{\overline{o}}(u(p))=\eta _{\mu \nu }ฯต_o^\nu (u(p))=u_\mu (p),`$ (C27)
$`ฯต_\mu ^r(u(p))=\delta ^{rs}\eta _{\mu \nu }ฯต_r^\nu (u(p))=(\delta ^{rs}u_s(p);\delta _j^r\delta ^{rs}\delta _{jh}{\displaystyle \frac{u^h(p)u_s(p)}{1+u^o(p)}}),`$ (C28)
$`ฯต_o^A(u(p))=u_A(p),`$ (C29)
which satisfy
$`ฯต_\mu ^A(u(p))ฯต_A^\nu (u(p))=\eta _\nu ^\mu ,`$ (C30)
$`ฯต_\mu ^A(u(p))ฯต_B^\mu (u(p))=\eta _B^A,`$ (C31)
$`\eta ^{\mu \nu }=ฯต_A^\mu (u(p))\eta ^{AB}ฯต_B^\nu (u(p))=u^\mu (p)u^\nu (p){\displaystyle \underset{r=1}{\overset{3}{}}}ฯต_r^\mu (u(p))ฯต_r^\nu (u(p)),`$ (C32)
$`\eta _{AB}=ฯต_A^\mu (u(p))\eta _{\mu \nu }ฯต_B^\nu (u(p)),`$ (C33)
$`p_\alpha {\displaystyle \frac{}{p_\alpha }}ฯต_A^\mu (u(p))=p_\alpha {\displaystyle \frac{}{p_\alpha }}ฯต_\mu ^A(u(p))=0.`$ (C34)
The Wigner rotation corresponding to the Lorentz transformation $`\mathrm{\Lambda }`$ is
$`R^\mu {}_{\nu }{}^{}(\mathrm{\Lambda },p)`$ $`=`$ $`[L(\stackrel{}{p},p)\mathrm{\Lambda }^1L(\mathrm{\Lambda }p,\stackrel{}{p})]^\mu {}_{\nu }{}^{}=\left(\begin{array}{cc}1& 0\\ 0& R^i{}_{j}{}^{}(\mathrm{\Lambda },p)\end{array}\right),`$ (C37)
$`R^i{}_{j}{}^{}(\mathrm{\Lambda },p)`$ $`=`$ $`(\mathrm{\Lambda }^1)^i{}_{j}{}^{}{\displaystyle \frac{(\mathrm{\Lambda }^1)^i{}_{o}{}^{}p_{\beta }^{}(\mathrm{\Lambda }^1)^\beta _j}{p_\rho (\mathrm{\Lambda }^1)^\rho {}_{o}{}^{}+\eta \sqrt{ฯตp^2}}}`$ (C39)
$``$ $`{\displaystyle \frac{p^i}{p^o+\eta \sqrt{ฯตp^2}}}[(\mathrm{\Lambda }^1)^o{}_{j}{}^{}{\displaystyle \frac{((\mathrm{\Lambda }^1)^o{}_{o}{}^{}1)p_\beta (\mathrm{\Lambda }^1)^\beta _j}{p_\rho (\mathrm{\Lambda }^1)^\rho {}_{o}{}^{}+\eta \sqrt{ฯตp^2}}}].`$ (C40)
The polarization vectors transform under the Poincarรฉ transformations $`(a,\mathrm{\Lambda })`$ in the following way
$$ฯต_r^\mu (u(\mathrm{\Lambda }p))=(R^1)_r{}_{}{}^{s}\mathrm{\Lambda }_{}^{\mu }{}_{\nu }{}^{}ฯต_{s}^{\nu }(u(p)).$$
(C41)
## D More on Dixonโs Multipoles.
Let us add other forms of the Dixon multipoles.
In the case of the fluid configurations treated in Section II and IV, the Hamilton equations generated by the Dirac Hamiltonian (166) in the gauge $`\stackrel{}{q}_{sys}0`$ \[$`\stackrel{}{\lambda }(\tau )=0`$\] imply \[in Ref. this is a consequence of $`_\mu T^{\mu \nu }\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$\]
$`{\displaystyle \frac{dp_T^\mu (T_s)}{dT_s}}`$ $`\stackrel{}{=}`$ $`0,forn=0,`$ (D1)
$`{\displaystyle \frac{dp_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)}{dT_s}}`$ $`\stackrel{}{=}`$ $`nu^{(\mu _1}(p_s)p_T^{\mu _2\mathrm{}\mu _n)\mu }(T_s)+nt_T^{(\mu _1\mathrm{}\mu _n)\mu }(T_s),n1.`$ (D2)
Let us define for $`n1`$
$`b_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ $`=`$ $`p_T^{(\mu _1\mathrm{}\mu _n\mu )}(T_s)=`$ (D3)
$`=`$ $`ฯต_{r_1}^{(\mu _1}(u(p_s))\mathrm{}.ฯต_{r_n}^{\mu _n}(u(p_s))ฯต_A^{\mu )}(u(p_s))I_T^{r_1..r_nA\tau }(T_s),`$ (D4)
$`ฯต_{\mu _1}^{r_1}(u(p_s))\mathrm{}.ฯต_{\mu _n}^{r_n}(u(p_s))b_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ $`=`$ $`{\displaystyle \frac{1}{n+1}}u^\mu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s)+ฯต_r^\mu (u(p_s))I_T^{(r_1\mathrm{}r_nr)\tau }(T_s),`$ (D6)
$`c_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ $`=`$ $`c_T^{(\mu _1\mathrm{}\mu _n)\mu }(T_s)=p_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)p_T^{(\mu _1\mathrm{}\mu _n\mu )}(T_s)=`$ (D8)
$`=`$ $`[ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^\mu (u(p_s))`$ (D9)
$``$ $`ฯต_{r_1}^{(\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))ฯต_A^{\mu )}(u(p_s))]I_T^{r_1..r_nA\tau }(T_s),`$ (D12)
$`c_T^{(\mu _1\mathrm{}\mu _n\mu )}(T_s)=0,`$
$`ฯต_{\mu _1}^{r_1}(u(p_s))\mathrm{}.ฯต_{\mu _n}^{r_n}(u(p_s))c_T^{\mu _1\mathrm{}\mu _n\mu }(T_s)`$ $`=`$ $`{\displaystyle \frac{n}{n+1}}u^\mu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s)+`$ (D13)
$`+`$ $`ฯต_r^\mu (u(p_s))[I_T^{r_1\mathrm{}r_nr\tau }(T_s)I_T^{(r_1\mathrm{}r_nr)\tau }(T_s)],`$ (D14)
and then for $`n2`$
$`d_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`=`$ $`d_T^{(\mu _1\mathrm{}\mu _n)(\mu \nu )}(T_s)=t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ (D15)
$``$ $`{\displaystyle \frac{n+1}{n}}[t_T^{(\mu _1\mathrm{}\mu _n\mu )\nu }(T_s)+t_T^{(\mu _1\mathrm{}\mu _n\nu )\mu }(T_s)]+`$ (D16)
$`+`$ $`{\displaystyle \frac{n+2}{n}}t_T^{(\mu _1\mathrm{}\mu _n\mu \nu )}(T_s)=`$ (D17)
$`=`$ $`[ฯต_{r_1}^{\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^\mu ฯต_B^\nu {\displaystyle \frac{n+1}{n}}(ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^{\mu )}ฯต_B^\nu +`$ (D18)
$`+`$ $`ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_B^{\nu )}ฯต_A^\mu )+{\displaystyle \frac{n+2}{n}}ฯต_{r_1}^{(\mu _1}..ฯต_{r_n}^{\mu _n}ฯต_A^\mu ฯต_B^{\nu )}](u(p_s))`$ (D22)
$`I_T^{r_1..r_nAB}(T_s),`$
$`d_T^{(\mu _1\mathrm{}\mu _n\mu )\nu }(T_s)=0,`$
$`ฯต_{\mu _1}^{r_1}(u(p_s))\mathrm{}.ฯต_{\mu _n}^{r_n}(u(p_s))d_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`=`$ $`{\displaystyle \frac{n1}{n+1}}u^\mu (p_s)u^\nu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s)+`$ (D24)
$`+`$ $`{\displaystyle \frac{1}{n}}[u^\mu (p_s)ฯต_r^\nu (u(p_s))+u^\nu (p_s)ฯต_r^\mu (u(p_s))]`$ (D26)
$`[(n1)I_T^{r_1\mathrm{}r_nr\tau }(T_s)+I_T^{(r_1\mathrm{}r_nr)\tau }(T_s)]+`$
$`+`$ $`ฯต_{s_1}^\mu (u(p_s))ฯต_{s_2}^\nu (u(p_s))[I_T^{r_1\mathrm{}r_ns_1s_2}(T_s)`$ (D27)
$``$ $`{\displaystyle \frac{n+1}{n}}(I_T^{(r_1\mathrm{}r_ns_1)s_2}(T_s)+I_T^{(r_1\mathrm{}r_ns_2)s_1}(T_s))+`$ (D28)
$`+`$ $`I_T^{(r_1\mathrm{}r_ns_1s_2)}(T_s)].`$ (D29)
Then Eqs.(D2) may be rewritten in the form
$`1)n=1`$ (D30)
$`t_T^{\mu \nu }(T_s)`$ $`=`$ $`t_T^{(\mu \nu )}(T_s)\stackrel{}{=}p_T^\mu (T_s)u^\nu (p_s)+{\displaystyle \frac{1}{2}}{\displaystyle \frac{d}{dT_s}}(S_T^{\mu \nu }(T_s)[\alpha ]+2b_T^{\mu \nu }(T_s)),`$ (D33)
$``$
$`t_T^{\mu \nu }(T_s)`$ $`\stackrel{}{=}`$ $`p_T^{(\mu }(T_s)u^{\nu )}(p_s)+{\displaystyle \frac{d}{dT_s}}b_T^{\mu \nu }(T_s)=P^\tau u^\mu (p_s)u^\nu (p_s)+P^ru^{(\mu }(p_s)ฯต_r^{\nu )}(u(p_s))+`$ (D34)
$`+`$ $`ฯต_r^{(\mu }(u(p_s))u^{\nu )}(p_s)I_T^{r\tau \tau }(T_s)+`$ (D35)
$`+`$ $`ฯต_r^{(\mu }(u(p_s))ฯต_s^{\nu )}(u(p_s))I_T^{rs\tau }(T_s),`$ (D36)
$`{\displaystyle \frac{d}{dT_s}}S_T^{\mu \nu }(T_s)[\alpha ]`$ $`\stackrel{}{=}`$ $`2p_T^{[\mu }(T_s)u^{\nu ]}(p_s)=2P_\varphi ^rฯต_r^{[\mu }(u(p_s))u^{\nu ]}(p_s)0,`$ (D39)
$`2)n=2[identityt_T^{\rho \mu \nu }=t_T^{(\rho \mu )\nu }+t_T^{(\rho \nu )\mu }+t_T^{(\mu \nu )\rho }]`$
$`2t_T^{(\rho \mu )\nu }(T_s)`$ $`\stackrel{}{=}`$ $`2u^{(\rho }(p_s)b_T^{\mu )\nu }(T_s)+u^{(\rho }(p_s)S_T^{\mu )\nu }(T_s)[\alpha ]+{\displaystyle \frac{d}{dT_s}}(b_T^{\rho \mu \nu }(T_s)+c_T^{\rho \mu \nu }(T_s)),`$ (D42)
$``$
$`t_T^{\rho \mu \nu }(T_s)`$ $`\stackrel{}{=}`$ $`u^\rho (p_s)b_T^{\mu \nu }(T_s)+S_T^{\rho (\mu }(T_s)[\alpha ]u^{\nu )}(p_s)+{\displaystyle \frac{d}{dT_s}}({\displaystyle \frac{1}{2}}b_T^{\rho \mu \nu }(T_s)c_T^{\rho \mu \nu }(T_s)),`$ (D45)
$`3)n3`$
$`t_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`\stackrel{}{=}`$ $`d_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)+u^{(\mu _1}(p_s)b_T^{\mu _2\mathrm{}\mu _n)\mu \nu }(T_s)+2u^{(\mu _1}(p_s)c_T^{\mu _2\mathrm{}\mu _n)(\mu \nu )}(T_s)+`$ (D47)
$`=`$ $`{\displaystyle \frac{2}{n}}c_T^{\mu _1\mathrm{}\mu _n(\mu }(T_s)u^{\nu )}(p_s)+{\displaystyle \frac{d}{dT_s}}[{\displaystyle \frac{1}{n+1}}b_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)+{\displaystyle \frac{2}{n}}c_T^{\mu _1\mathrm{}\mu _n(\mu \nu )}(T_s)],`$ (D48)
This allows to rewrite $`<T^{\mu \nu },f>`$ in the following form
$`<T^{\mu \nu },f>`$ $`=`$ $`{\displaystyle }dT_s{\displaystyle }{\displaystyle \frac{d^4k}{(2\pi )^4}}\stackrel{~}{f}(k)e^{ikx_s(T_s)}[u^{(\mu }(p_s)p_T^{\nu )}(T_s)ik_\rho S_T^{\rho (\mu }(T_s)[\alpha ]u^{\nu )}(p_s)+`$ (D49)
$`+`$ $`{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(i)^n}{n!}}k_{\rho _1}\mathrm{}k_{\rho _n}_T^{\rho _1\mathrm{}\rho _n\mu \nu }(T_s)],`$ (D50)
with
$`_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`=`$ $`_T^{(\mu _1\mathrm{}\mu _n)(\mu \nu )}(T_s)=d_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ (D51)
$``$ $`{\displaystyle \frac{2}{n1}}u^{(\mu _1}(p_s)c_T^{\mu _2\mathrm{}\mu _n)(\mu \nu )}(T_s)+{\displaystyle \frac{2}{n}}c_T^{\mu _1\mathrm{}\mu _n(\mu }(T_s)u^{\nu )}(p_s)=`$ (D52)
$`=`$ $`[ฯต_{r_1}^{\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^\mu ฯต_B^\nu {\displaystyle \frac{n+1}{n}}(ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^{\mu )}ฯต_B^\nu +`$ (D53)
$`+`$ $`ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_B^{\nu )}ฯต_A^\mu )+{\displaystyle \frac{n+2}{n}}ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^\mu ฯต_B^{\nu )}](u(p_s)0I_T^{r_1..r_nAB}(T_s)`$ (D54)
$``$ $`[{\displaystyle \frac{2}{n1}}u^{(\mu _1}(p_s)(ฯต_{r_1}^{\mu _2}\mathrm{}ฯต_{r_{n1}}^{\mu _n)}ฯต_{r_n}^{(\mu }ฯต_A^{\nu )}ฯต_{r_1}^{(\mu _2}\mathrm{}ฯต_{r_{n1}}^{\mu _n)}ฯต_{r_n}^{(\mu }ฯต_A^{\nu ))})`$ (D55)
$``$ $`{\displaystyle \frac{2}{n}}(ฯต_{r_1}^{\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^{(\mu }ฯต_{r_1}^{(\mu _1}\mathrm{}ฯต_{r_n}^{\mu _n}ฯต_A^{(\mu )}u^{\nu )}(p_s)](u(p_s)0I_T^{r_1..r_nA\tau }(T_s),`$ (D58)
$`_T^{(\mu _1\mathrm{}\mu _n\mu )\nu }(T_s)=0,`$
$`ฯต_{\mu _1}^{r_1}(u(p_s))\mathrm{}.ฯต_{\mu _n}^{r_n}(u(p_s))_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`=`$ $`{\displaystyle \frac{n+3}{n+1}}u^\mu (p_s)u^\nu (p_s)I_T^{r_1\mathrm{}r_n\tau \tau }(T_s)+`$ (D60)
$`+`$ $`{\displaystyle \frac{1}{n}}[u^\mu (p_s)ฯต_r^\nu (u(p_s))+u^\nu (p_s)ฯต_r^\mu (u(p_s))]I_T^{r_1\mathrm{}r_nr\tau }(T_s)+`$ (D61)
$`+`$ $`ฯต_{s_1}^\mu (u(p_s))ฯต_{s_2}^\nu (u(p_s))[I_T^{r_1\mathrm{}r_ns_1s_2}(T_s)`$ (D62)
$``$ $`{\displaystyle \frac{n+1}{n}}(I_T^{(r_1\mathrm{}r_ns_1)s_2}(T_s)+I_T^{(r_1\mathrm{}r_ns_2)s_1}(T_s))+`$ (D63)
$`+`$ $`I_T^{(r_1\mathrm{}r_ns_1s_2)}(T_s)].`$ (D64)
Finally, a set of multipoles equivalent to the $`_T^{\mu _1\mathrm{}\mu _n\mu \nu }`$ is
$`n0`$ (D65)
$`J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }(T_s)`$ $`=`$ $`J_T^{(\mu _1\mathrm{}\mu _n)[\mu \nu ][\rho \sigma ]}(T_s)=_T^{\mu _1\mathrm{}\mu _n[\mu [\rho \nu ]\sigma ]}(T_s)=`$ (D67)
$`=`$ $`t_T^{\mu _1\mathrm{}\mu _n[\mu [\rho \nu ]\sigma ]}(T_s){\displaystyle \frac{1}{n+1}}[u^{[\mu }(p_s)p_T^{\nu ]\mu _1\mathrm{}\mu _n[\rho \sigma ]}(T_s)+`$ (D68)
$`+`$ $`u^{[\rho }(p_s)p_T^{\sigma ]\mu _1\mathrm{}\mu _n[\mu \nu ]}(T_s)]=`$ (D69)
$`=`$ $`[ฯต_{r_1}^{\mu _1}..ฯต_{r_n}^{\mu _n}ฯต_r^{[\mu }ฯต_s^{[\rho }ฯต_A^{\nu ]}ฯต_B^{\sigma ]}](u(p_s))I_T^{r_1..r_nAB}(T_s)`$ (D70)
$``$ $`{\displaystyle \frac{1}{n+1}}[u^{[\mu }(p_s)ฯต_r^{\nu ]}(u(p_s))ฯต_s^{[\rho }(u(p_s))ฯต_A^{\sigma ]}(u(p_s))+`$ (D71)
$`+`$ $`u^{[\rho }(p_s)ฯต_r^{\sigma ]}(u(p_s))ฯต_s^{[\mu }(u(p_s))ฯต_A^{\nu ]}(u(p_s))]`$ (D77)
$`ฯต_{r_1}^{\mu _1}(u(p_s))\mathrm{}ฯต_{r_n}^{\mu _n}(u(p_s))I_T^{rr_1..r_nsA\tau }(T_s),`$
$`[(n+4)(3n+5)linearlyindependentcomponents],`$
$`n1`$
$`u_{\mu _1}(p_s)`$ $`J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }(T_s)=J_T^{\mu _1\mathrm{}\mu _{n1}(\mu _n\mu \nu )\rho \sigma }(T_s)=0,`$ (D81)
$`n2`$
$`_T^{\mu _1\mathrm{}\mu _n\mu \nu }(T_s)`$ $`=`$ $`{\displaystyle \frac{4(n1)}{n+1}}J_T^{(\mu _1\mathrm{}\mu _{n1}|\mu |\mu _n)\nu }(T_s),`$ (D83)
$`ฯต_{\mu _1}^{r_1}(u(p_s))\mathrm{}.ฯต_{\mu _n}^{r_n}(u(p_s))J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }(T_s)`$ $`=`$ $`\left[ฯต_r^{[\mu }ฯต_s^{[\rho }ฯต_A^{\nu ]}ฯต_B^{\sigma ]}\right](u(p_s))I_T^{r_1..r_nAB}(T_s)`$ (D85)
$``$ $`{\displaystyle \frac{1}{n+1}}[u^{[\mu }(p_s)ฯต_r^{\nu ]}(u(p_s))ฯต_s^{[\rho }(u(p_s))ฯต_A^{\sigma ]}(u(p_s))+`$ (D86)
$`+`$ $`u^{[\rho }(p_s)ฯต_r^{\sigma ]}(u(p_s))ฯต_s^{[\mu }(u(p_s))ฯต_A^{\nu ]}(u(p_s))]I_T^{rr_1..r_nsA\tau }(T_s).`$ (D87)
The $`J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }`$ are the Dixon โ$`2^{n+2}`$-pole inertial moment tensorsโ of the extended system: they \[or equivalently the $`_T^{\mu _1\mathrm{}\mu _n\mu \nu }`$โs\] determine its energy-momentum tensor together with the monopole $`p_T^\mu `$ and the spin dipole $`S_T^{\mu \nu }`$. The equations $`_\mu T^{\mu \nu }\stackrel{}{=}\mathrm{\hspace{0.17em}0}`$ are satisfied due to the equations of motion (D48) for $`P_T^\mu `$ and $`S_T^{\mu \nu }`$ \[the so called Papapetrou-Dixon-Souriau equations given in Eqs.(381)\] without the need of the equations of motion for the $`J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }`$. When all the multipoles $`J_T^{\mu _1\mathrm{}\mu _n\mu \nu \rho \sigma }`$ are zero \[or negligible\] one speaks of a pole-dipole field configuration of the perfect fluid. |
warning/0003/astro-ph0003486.html | ar5iv | text | # HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede
## 1 Introduction
Recent observations from both the Galileo spacecraft and the Hubble Space Telescope (HST) have considerably altered our knowledge of the atmospheres of the Jovian satellites Europa and Ganymede. Both are now known to have tenuous atmospheres (column density $`5\times 10^{14}`$ cm<sup>-2</sup>) with both molecular oxygen (Europa and Ganymede) and atomic hydrogen (Ganymede) components. The Galileo UV spectrometer detected H i Lyman-$`\alpha `$ emission at Ganymede from a hydrogen exosphere (Barth et al., 1997), and charged particle measurements indicated that there is also an outflow of protons, implying ongoing gas production (Frank et al., 1997). The oxygen component was detected through HST/Goddard High Resolution Spectrograph (GHRS) observations of the atomic oxygen multiplets O i $`\lambda `$1304 and O i\] $`\lambda `$1356 (Hall et al., 1995, 1998). The intensity ratio of these emissions implies that the primary source is electron dissociative excitation of molecular oxygen. The source of both the hydrogen and O<sub>2</sub> is thought to be sputtering of surface water ice by Io plasma torus ions. Beyond the fact that they exist, very little is known about these atmospheres, including their vertical structure, areal coverage, and variability, which could be significant if the dominant source is surface sputtering because of the asymmetric nature of the plasma bombardment. Plasma bombardment of the surface is also supported by the recent detection of ozone and O<sub>2</sub> embedded in the surface ice of Ganymede, and SO<sub>2</sub> embedded in the surface ice of Europa (Spencer et al., 1995; Calvin et al., 1996; Noll et al., 1995, 1996).
Galileo magnetometer measurements have also shown strong perturbations in the Jovian magnetic field near Ganymede (Kivelson et al., 1996, 1997). The measured perturbations indicate that the satellite possesses a magnetic field sufficiently strong ($``$1500 nT) to overpower Jupiterโs ambient field, and that Ganymedeโs magnetic and spin axes are roughly aligned (Kivelson et al., 1996). Near Ganymede closest approach the plasma wave experiment also detected a significant population of trapped, charged particles (Gurnett et al., 1996), implying that Ganymede possesses a โmagnetosphere within a magnetosphere.โ Consistent with these results, the HST/GHRS observations (Hall et al., 1998) have raised the intriguing possibility that Ganymede exhibits polar aurora. In those spectra the Ganymede O i\] $`\lambda `$1356 emission line exhibits a doubly-peaked profile that is inconsistent with that of a diffuse source filling the aperture, or with emission from a uniform disk. Hall et al. postulated that the double-peaked profile implies the existence of a similarly double-peaked structure in the spatial distribution of the emission source within the aperture, with the strongest emissions coincident with Ganymedeโs north and south polar regions. In this paper, we report ultraviolet objective grating images of Ganymede made with the Space Telescope Imaging Spectrograph (STIS) (installed in HST in February 1997) which confirm the spatial distribution inferred from the earlier observation and raise new questions about the interaction of Ganymedeโs atmosphere with the Jovian magnetosphere.
## 2 Observations
Observations were obtained over four contiguous HST orbits on 1998 October 30 with the STIS G140L grating using the $`2^{\prime \prime }\times 25^{\prime \prime }`$ slit. At the time of observation, Ganymede was 4.25 AU from Earth, its sub-Earth longitude varied from 290 to 300 and the phase angle was 8.6. Since the diameter of Ganymedeโs disk was $`1.^{\prime \prime }71`$, this provided effective objective grating spectroscopy over the wavelength range of 1160 โ 1720 ร
. A log of the exposures is given in Table 1. The STIS mode used for the observations is the same as that used to observe Io, and the details can be found in Roesler et al. (1999). Two distinct time-tagged spectral images are obtained in each of four HST orbits. The first image in orbits 2โ4 exhibits a very high geocoronal Lyman-$`\alpha `$ background (typically 15 kR as opposed to 3.5 kR during the dark part of an HST orbit) as well as strong O i $`\lambda `$1304 airglow emission due to the illumination of the Earthโs upper atmosphere by sunlight. An example of a raw spectral image is shown in Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede. Note the $`2^{\prime \prime }`$ wide vertical stripe corresponding to geocoronal Lyman-$`\alpha `$ with the disk reflected Lyman-$`\alpha `$ image of Ganymede perfectly centered in in the slit. The raw image also shows the two oxygen multiplets, clearly separated, and reflected sunlight at the longer wavelengths.
## 3 Discussion
### 3.1 Spectra
Individual spectra (two per orbit) were extracted by summing the data over 82 pixels ($`2^{\prime \prime }`$) along the slit centered on Ganymede. The background, particularly the geocoronal Lyman-$`\alpha `$, was obtained by summing 151 pixels ($`3.^{\prime \prime }68`$) along the slit on both sides of Ganymede, and averaging the two. The results for a single spectral image, with the background subtracted, are shown in Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede. In this figure, the one-dimensional extracted spectrum has also been rebinned by four pixels to enhance the signal/noise ratio. The unusual shape of the spectral lines of H i Lyman-$`\alpha `$ and the two O i emissions result from the spatial distribution of these emissions on the disk of Ganymede.
Longward of 1380 ร
the signal is reflected solar radiation. To model this component, a solar spectrum taken with the SOLSTICE instrument on UARS (Woods et al., 1996) appropriate to the level of solar activity in October 1998 was convolved with an assumed uniform reflecting disk of Ganymedeโs radius (note that one pixel is $`0.^{\prime \prime }0244\times 0.^{\prime \prime }0244`$ and the dispersion is 0.584 ร
pixel<sup>-1</sup>). This is overplotted in Fig. HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede. From this fit, the planetary albedo can be derived from the ratio of the reflected flux to the solar flux at Jupiter:
$$p(\lambda )=\frac{F_G(\lambda )\pi d^2}{F_{}(\lambda )\mathrm{\Omega }_G}\varphi (\theta ,\lambda )$$
where $`d`$ is the Sun-Jupiter distance (in AU), $`\mathrm{\Omega }_G`$ is the solid angle of Ganymede as seen from Earth, $`F_{}(\lambda )`$ is the solar flux at 1 AU, and $`\varphi (\theta ,\lambda )`$ is the phase function at phase angle $`\theta `$. To determine the albedo at a wavelength as close as possible to the oxygen emissions, we choose a 50 ร
band centered at 1405 ร
. With the assumption of unity for $`\varphi (\theta ,\lambda )`$, the derived albedo near 1400 ร
is found to be $`2.3\pm 0.2`$%, in good agreement with the value of $`2.6\pm 0.3`$% derived by Hall et al. (1998) from GHRS measurements of the reflected C ii $`\lambda `$1335 multiplet. Note that the measurement of Hall et al. was made at a slightly smaller phase angle, 2.7.
The lower panel of Fig. HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede shows a single spectrum after subtraction of the fitted solar spectrum assuming a constant albedo with wavelength. The two O i emissions are clearly separated and both have a shape determined by the spatial distribution on the disk, similar to that inferred by Hall et al. (1998) from their one-dimensional spectra. The fluxes of the two multiplets are extracted and tabulated, together with the ratio of the two (O i\] $`\lambda `$1356/O i $`\lambda `$1304), in Table 1. The values are generally consistent with those reported by Hall et al. (note that Hall et al. give the total O i $`\lambda `$1304 flux, airglow plus reflected solar radiation), but the orbit to orbit variation suggests a real variability, one that is correlated with the changes in the morphology of the emissions discussed in the next section. The ratio of O i\] $`\lambda `$1356 to O i $`\lambda `$1304 is, like that found by Hall et al., consistently lower (but within the $`3\sigma `$ uncertainty) than the values expected for electron impact excitation of O<sub>2</sub> alone, 1.6โ2.0, indicating a possible contribution from electron impact of atomic oxygen to the O i $`\lambda `$1304 emission. However, this is not quantifiable as there is an indication, from Fig. HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede, that the albedo is increasing below 1300 ร
in which case the O i $`\lambda `$1304 flux would be over-estimated and the true intensity ratio would be closer to the known value for O<sub>2</sub> excitation.
### 3.2 O i\] $`\lambda `$1356 Images
Images of O i\] $`\lambda `$1356 were constructed using the flat-fielded counts (โfltโ) files from the HST pipeline rather than the fluxed two-dimensional image (โx2dโ) files used to generate Fig. HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede. This was done to avoid distortion introduced by the changing sensitivity across the $`2^{\prime \prime }`$ wide slit, which spans 47 ร
. To allow for temporal variability, the two separate exposures from each orbit were added together since the O i\] $`\lambda `$1356 emission is not affected by the higher Lyman-$`\alpha `$ or O i $`\lambda `$1304 background levels. Detector background was evaluated away from Ganymede along the slit and subtracted from the resulting $`82\times 82`$ pixel array. Each image was then rotated to align Jovian north along the vertical axis, rebinned to $`41\times 41`$ pixels (each pixel now $`0.^{\prime \prime }049`$ on a side), and smoothed by 3 in both directions. The results are shown in Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede, together with brightness contours calibrated in rayleighs. The images are characterized by bright polar regions and appear to be variable with time.
The location of the emissions, mostly at geographic latitudes above $`|40|`$ in both hemispheres, is in agreement with the model of Kivelson et al. (1997) that predicts the presence of Jovian field lines linked to Ganymede only at high latitudes. In addition, the regions of brightest emission occur at the geomagnetic latitudes where the separatrix regions intersect the atmosphere and which define the boundaries of the polar caps (Neubauer, 1998). Magnetic field line reconnection occurs along the separatrix imparting an atmospheric signature of enhanced conductivity and current known as the auroral or polar electrojets. Over the course of the four orbits the orientation of the Jovian magnetic field relative to Ganymedeโs magnetic field, which is tilted 10 from the spin axis (Kivelson et al., 1997), varied considerably, as indicated in Table 1. Thus, the locations of the polar caps and separatrix regions on the surface of Ganymede vary considerably over a Jovian rotation. This can account for both the variations in total O i flux and the relative brightnesses of the northern and southern hemisphere emissions.
However, there are two surprising aspects to these images, because Jovian magnetospheric electrons have direct access to the polar cap atmospheres, as implied by particle and fields measurements on Galileo (Williams et al., 1998; Paranicas et al., 1999). They are the longitudinal non-uniformity in the emission brightness at high latitudes, particularly in the southern hemisphere, and the lack of pronounced limb brightening near the poles, even at the smoothed 200 km spatial resolution of our images. The images yield $`50`$ R limb intensity above the polar caps. Under uniform conditions this is equivalent to $`15`$ R at 60 latitude and below our detection limit. The brightest regions on the disk are 20 times the disk intensity constrained by the observed limb intensity.
To explore the significance of the observed oxygen brightness, a model atmosphere was constructed with surface O<sub>2</sub> density of $`1\times 10^8`$ cm<sup>-3</sup> and column density of $`5.2\times 10^{14}`$ cm<sup>-2</sup>, which is consistent with the range adopted by Hall et al. (1998) and at the abundance upper limit deduced from a UV stellar occultation observed by Voyager (Broadfoot et al., 1981). For the electron density, a model was generated from measurements by the Galileo plasma wave instrument along fly-by trajectories and extrapolated to the surface with density of 370 cm<sup>-3</sup> (Gurnett et al., 1996). Unfortunately, there are no observational constraints on the electron temperature. From the data of Sittler & Strobel (1987) at their maximum L shell of 13, one would expect the Jovian magnetospheric electron temperature to be at least 20 eV. For a $`9`$ eV multiplet, the excitation rate would be insensitive to this or higher electron temperatures. With the above assumptions at $`T_e20`$ eV applied to the polar atmosphere on open field lines, electron impact dissociative excitation of O<sub>2</sub> yields 300 R of O i\] $`\lambda `$1356 at high latitudes with limb brightening to $`1`$ kR. A polar limb intensity of 50 R implies an O<sub>2</sub> column density of $`3\times 10^{13}`$ cm<sup>-2</sup>. Alternatively, in the absence of a direct measurement of Jovian electron temperature, the upper limit column density of $`5.2\times 10^{14}`$ cm<sup>-2</sup> is permissible, if $`T_e4`$ eV. Also possible would be various combinations of O<sub>2</sub> column density in the range of $`(0.35)\times 10^{14}`$ cm<sup>-2</sup> and electron temperature in the range of 4โ20 eV or higher.
There are additional factors to consider in the interpretation of the HST images. There is no evidence that Ganymedeโs surface temperature drops to the O<sub>2</sub> condensation temperature of 80 K which could account for an inhomogeneous atmosphere. In the range $`T_e=1100`$ eV, the calculated intensity ratio of O i\] $`\lambda `$1356 to O i$`\lambda `$1304 is limited to 1.6โ2.0 for a pure O<sub>2</sub> atmosphere. Lower values require the addition of atomic oxygen. In the limit of a pure atomic oxygen atmosphere, this intensity ratio has a value of 1.2 at $`T_e=4`$ eV, monotonically decreasing to $`0.35`$ at 20 eV. There is no apparent correlation of this ratio in Table 1 with absolute O i\] $`\lambda `$1356 brightness. Clearly to achieve the higher observed ratios requires an O<sub>2</sub> atmosphere. Finally, Jovian magnetospheric electrons on open field lines do not have large density variability on the length scales characteristic of the brightness variations in Fig. HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede (Gurnett et al., 1996).
Thus, the best explanation for the inhomogeneity in the emission brightness is that it is truly auroral in nature, analogous with the Earthโs highly variable UV luminosity in the auroral oval regions and driven by acceleration processes of electrons trapped within Ganymedeโs magnetosphere near the separatrix regions. Without knowledge of the distribution function of these auroral precipitating electrons and associated fluxes, the column density of Ganymedeโs atmosphere cannot be inferred from the bright auroral regions in HST images. For the present time the polar limb intensities are the only constraint on the atmospheric column density and until the temperature of electrons on open field lines is determined, this constraint is not firm.
### 3.3 H i Lyman-$`\alpha `$
Barth et al. (1997) have reported the detection of H i Lyman-$`\alpha `$ emission above the limb of Ganymede extending nearly one Ganymede radius (2634 km), which they attribute to a hydrogen exosphere. Such emission should be detectable in our long-slit spectral image but is masked by the strong geocoronal Lyman-$`\alpha `$ emission that fills the entire $`2^{\prime \prime }`$ wide slit (see Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede). To remove the geocoronal component, a Lyman-$`\alpha `$ โflat fieldโ along the slit is needed. This is obtained from our data in the following manner. The final three orbits contain separate spectral images with distinctly different values of geocoronal background, 15 kR for the first of each pair, 3.5 kR for the second. The three โhighโ images and the three โlowโ images are separately combined (again using the flat-fielded counts files rather than the flux calibrated files), and spatial profiles along the slit (summing 82 pixels in the dispersion direction) are obtained. These are shown in the top panel of Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede. The profiles are normalized to the slightly different cumulative exposure times and the difference is taken, which eliminates the signal due to Ganymede, and this is also shown (after median filtering) in the figure. The geocoronal background is then normalized to and subtracted from the โlowโ image giving the net Lyman-$`\alpha `$ spatial profile associated with Ganymede, as shown in the lower panel of Figure HST/STIS Ultraviolet Imaging of Polar Aurora on Ganymede, where emission above both limbs is clearly detected. The radial model of Barth et al. (1997), integrated across the width of the STIS slit, is also shown in the figure and is found to fit our data very well.
## 4 Conclusions
Objective grating images of Ganymede obtained with HST/STIS show clearly separated O i emissions confirming the result of Hall et al. (1998) that the emissions are confined to polar regions (latitudes above 45). The total fluxes are consistent with those reported by Hall et al. but appear to vary in time and in the relative intensities between northern and southern hemispheres. The O i\] $`\lambda `$1356/O i $`\lambda `$1304 ratio is consistent with the primary excitation mechanism being electron impact on O<sub>2</sub>, as postulated by Hall et al. While the spatial distribution of the emissions is consistent with current models of the magnetic field of Ganymede, expected longitudinal uniformity and limb brightening are not observed. In addition, Lyman-$`\alpha `$ limb emission from a hydrogen exosphere is detected and the measured brightness is found to be in good agreement with the Galileo UVS observations of Barth et al. (1997).
This work is based on observations with the National Aeronautics and Space Administration โ European Space Agency HST obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. We acknowledge partial support by NASA contract NAS5-30403 to the Johns Hopkins University. |
warning/0003/cond-mat0003394.html | ar5iv | text | # Model of Hydrophobic Attraction in Two and Three Dimensions
## I Introduction
A one-dimensional lattice model of hydrophobic attraction has recently been studied . The model incorporates what is believed to be the basic mechanism of the hydrophobic effect; viz., that the accommodation of the solute in the solvent is energetically favorable but sufficiently unfavorable entropically as to result in an increase in free energy. This increase is smaller when the hydrophobes are close together than when they are widely separated, resulting in a net solvent-mediated attraction between them.
The unfavorable entropy of accommodating the solute is achieved in the model by allowing each solvent molecule to have a large number $`q`$ of possible orientations but requiring two neighboring solvent molecules to be both in a special one of those $`q`$ states if they are to accommodate a hydrophobic solute in the interstitial site between them. That circumstance in which two neighboring solvent molecules are both in that special orientation is taken to be energetically favorable by an amount previously called $`uw`$. The condition of hydrophobicity at temperature $`T`$ was that $`kT\mathrm{ln}(q1)>uw`$, with $`k`$ Boltzmannโs constant. To model the hydrophobic effect realistically, $`(uw)/k`$ was taken to be $`3000`$K and $`q`$ was taken to be 110,000.
It was then a property of the model that forcing the solvent to accommodate a hydrophobic solute greatly restricted (to one state out of $`q`$) the possible orientations of the solvent molecules neighboring the accommodated solute. Such restriction of solvent orientation by the solute is consistent with experimental measurements of NMR relaxation times, which show that the reorientation times of water molecules that neighbor a solute with alkyl groups are about twice as long as in pure water, and that there is an increased activation energy for the reorientation in the presence of the solute over that in pure water.
It was a striking feature of the one-dimensional model that the range and strength of the solvent-mediated attraction between solutes were inversely related to each other; at low temperatures the attraction was weak but long-ranged and at high temperatures it was strong but short-ranged. One of the motivations of the present work was to see if that inverse relation, which is of potential importance in the interpretation of experiment, persists in higher dimensions, and it is indeed found to be so.
It was also found earlier in the one-dimensional model that the solubility of the hydrophobe decreases with increasing temperature, which is consistent with the increasing hydrophobicity that manifests itself in the increasing strength of the hydrophobic attraction. This temperature dependence of the solubility, too, is found here to persist in the two- and three-dimensional models. It is also found, as in the one-dimensional model, that if the energy of interaction of the hydrophobic solute with its solvent neighbors (earlier called $`v`$) is greater than the magnitude $`uw>0`$ of the favorable energy of solvent alignment, then the solubility, after first decreasing with increasing temperature, reaches a minimum and then increases.
The two- and three-dimensional versions of the model are defined in the following section. Also in Section II are the basic formulas that relate the solvent-mediated part of the potential of mean force and the solubility to the quantities that are measured in the simulations. The details of the simulations are in Section III, where the results are also displayed. These are discussed and summarized in the concluding Section IV.
## II The two- and three-dimensional models
The two-dimensional model is pictured in Fig. 1. At each site of a square lattice there is a solvent molecule that may be in any of a large number, $`q`$, of different states (orientations). Solute molecules may only be accommodated at the interstitial sites between neighboring pairs of solvent molecules, but only if that site is not already occupied and then only if the two solvent molecules in question are both in a special one of the $`q`$ states, called state 1 in what follows. The figure shows three solute molecules so accommodated.
The three-dimensional model is defined similarly, and also on a lattice of coordination number 4, as in two dimensions. This three-dimensional lattice is pictured in Fig. 2. It is obtained by systematically removing two bonds from each vertex of a simple cubic lattice, in such a way that from each site $`(i,j,k)`$ there remains a bond to the neighboring sites at $`(i+1,j,k)`$ and $`(i1,j,k)`$, as well as to those at $`(i,j1,k)`$ and $`(i,j,k1)`$ if $`i+j+k`$ is even but to those at $`(i,j+1,k)`$ and $`(i,j,k+1)`$ instead if $`i+j+k`$ is odd. The lattice is topologically (not metrically) equivalent to the tetrahedral diamond lattice, which is also the lattice of oxygen atoms in the metastable phase of ice called ice I<sub>c</sub>.
In both the two- and three-dimensional versions of the model only neighboring solvent molecules interact with each other, which they do with energy $`w`$ if both are in the special state 1 and with energy $`u>w`$ otherwise. The only solvent molecules with which a solute molecule interacts are its immediate neighbors; the sum of its energies of interaction with them is $`v`$.
Solute molecules may interact with each other but that is irrelevant here; the solute-solute interaction does not enter the expression for the solvent-mediated part of the potential of mean force between pairs of solutes in the infinitely dilute solution , nor is it relevant for the solubility of the solute, which is so slight that solute-solute interactions are negligible. The solute-solvent interaction parameter $`v`$ is also absent from the expression for the potential of mean force in the infinitely dilute solution but it is a significant parameter in the solubility.
Because the coordination number of the lattices considered here is 4, rather than 2 as in one dimension, the values of the parameters $`uw`$ and $`q`$ required to model the hydrophobic effect conveniently but still realistically over the temperature range 275-375K are both much smaller than in one dimension. Here, in both the two- and three-dimensional models, they are taken to be $`(uw)/k=1000`$K instead of 3000K and $`q=1100`$ instead of 110,000. It is still the case, as in the one-dimensional model, that $`kT\mathrm{ln}(q1)>uw`$ over the relevant temperature range, and that the extent to which $`kT\mathrm{ln}(q1)`$ exceeds $`uw`$ increases with increasing temperature โ which are the respective conditions required for the hydrophobic effect to manifest itself and to do so with increasing strength as the temperature increases.
Let $`p_{11}`$ be the probability that in the pure solvent both molecules of a given neighboring pair will be found in the special state 1; and let $`p(r)`$ be the probability that the molecules at two such pairs of sites will all be found to be in the special state 1 when the corresponding two interstitial sites, one associated with each pair, are separated by the metrical distance $`r`$ (measured in units of the lattice spacing). In the two-dimensional lattice of Fig. 1 the values of $`r`$ that occur are $`r=\sqrt{2}/2=0.71;1;\sqrt{2}=1.41;\sqrt{5/2}=1.58`$; etc. In the three-dimensional lattice of Fig. 2 the values of $`r`$ that occur are those and also $`r=\sqrt{3/2}=1.22`$, etc.
Let $`W(r)`$ be the solvent-mediated part of the potential of mean force between two solute molecules a distance $`r`$ apart, in the limit of infinite dilution. This is the potential of mean force from which the direct interaction between the two solute molecules themselves has been subtracted, and is the part of the mean-force potential that is of interest here. Then, as in the one-dimensional model, $`W(r)`$ is given in terms of $`p_{11}`$ and $`p(r)`$ by
$$W(r)=kT\mathrm{ln}\left[p(r)/p_{11}^2\right].$$
(1)
The quantities $`p_{11}`$ and $`p(r)`$ are what are determined in the simulations and $`W(r)`$ is then obtained from (1). It is a property of the solvent alone, and so depends only on the parameters $`uw`$ and $`q`$ but not on $`v`$.
The solubility is defined as in the earlier one-dimensional model. One imagines a hypothetical ideal gas of pure solute, of number density $`\rho _g`$, in osmotic equilibrium with the saturated solution; i.e., having a thermodynamic activity equal to that of the solute in the solution. Then if $`\rho _s`$ is the number density of the solute in the solution, the solubility $`\mathrm{\Sigma }`$ is here defined as the dimensionless ratio
$$\mathrm{\Sigma }=\rho _s/\rho _g.$$
(2)
If the solubility is low so that the saturated solution is very dilute, this $`\mathrm{\Sigma }`$ is the Ostwald absorption coefficient and is given by
$$\mathrm{\Sigma }=p_{11}\mathrm{e}^{v/kT}.$$
(3)
Thus, $`\mathrm{\Sigma }`$ is obtained once $`p_{11}`$ is calculated and a value of the solute-solvent interaction energy $`v`$ is specified.
## III Simulations
Properties of the two- and three-dimensional models presented in the previous Section are explored by means of Monte Carlo simulations. In these simulations, one solvent molecule is placed on each lattice site. For each molecule, we keep track of whether its state is the special state 1, or not. Solute molecules are not present in the Monte Carlo simulations, since the aim of the simulations is to numerically estimate the quantities $`p_{11}`$ and $`p(r)`$, from which the solvent-mediated part of the potential of mean force can be computed using (1).
The dynamics of the models are not specified, and we are therefore free to choose any dynamics that yield the correct equilibrium properties. In our Monte Carlo simulations, the dynamics consist of two processes: 1) each solvent molecule that is in its special state 1 leaves this special state with a rate of $`q1`$ per time unit; 2) each solvent molecule that is in a different state enters its special state with a rate of $`\alpha ^n`$ per time unit, where $`\alpha =\mathrm{exp}((uw)/kT)`$ and $`n`$ is the number of nearest-neighbor sites that are occupied with a solvent molecule in its special state. An efficient implementation of these dynamics is obtained with a BKL-scheme, after Bortz, Kalos and Lebowitz .
The simulations are performed on a $`500\times 500`$ lattice for the two-dimensional model, and a $`50\times 50\times 50`$ lattice for the three-dimensional model, both with periodic boundary conditions in order to reduce finite-size effects. In the initial configuration, no molecule is in its special state. This configuration is then thermalized for 0.1 time units. After thermalization, $`p_{11}`$ and $`p(r)`$ are measured every $`10^3`$ time unit, up to a total time of $`10^3`$ time units, and averaged. In a simulation of this length and for these system sizes, the two-dimensional and three-dimensional model undergo roughly $`5.510^8`$ and $`2.710^8`$ local changes, respectively. This procedure is repeated 10 times for each of the temperatures $`T`$=275, 300, 325, 350, and 375K, and for the three-dimensional model also at a temperature of 250K at which the system is probably metastable, the stable regime being one where almost all atoms are in their special state.
The values for the quantity $`p_{11}`$ are presented in Table I, for the two- and three-dimensional models. This quantity is decreasing with increasing temperature. The quantity $`p_{11}`$ is one of the two factors determining the solubility as defined in equation (3), the other being $`\mathrm{exp}(v/kT)`$. While this latter factor is monotonically increasing with increasing temperature for $`v>0`$, our data show that $`p_{11}`$ is monotonically decreasing with increasing temperature. For some choices of $`v`$, the combination of these two factors leads to a solubility that is initially decreasing with increasing temperature (a signature of the hydrophobic effect), but later increasing.
The solvent-mediated part of the potential of mean force $`W(r)`$ is plotted as a function of the distance $`r`$ in Fig. 3 for the two-dimensional model, and in Fig. 4 for the three-dimensional model. In the latter figure, the curve obtained at a temperature of 250K is dashed, to indicate its likely metastability. In agreement with the one-dimensional model, the strength of the hydrophobic interaction increases with increasing temperature, while its range decreases at the same time: an inverse relation is observed between the strength and range of the hydrophobic interaction, althoug less prominently in three dimensions than in one or two. Unlike the one-dimensional model, the two- and three-dimensional models show oscillations in the mean force, reflecting the greater geometrical complexity of the lattice in higher dimensions. In that sense, the curves resemble more closely what is seen in models in which the solvent is modeled more realistically . In addition to the strong attraction at short distances, there is now also a local minimum in the potential of depth about $`kT`$ at a distance of three lattice spacings in two dimensions and one of depth about $`2kT`$ at a distance of two lattice spacings in three dimensions.
## IV Summary and conclusions
A one-dimensional lattice model of hydrophobic attraction that was studied recently is extended to two and three dimensions. Monte Carlo simulations have been used to determine the solvent-mediated contribution to the potential of mean force between hydrophobic solute molecules, and the solubility of the solute.
As in the one-dimensional model, an inverse relation is observed between the strength and range of the hydrophobic interaction. With increasing temperature, the strength of the hydrophobic interaction increases, while at the same time its range decreases, although the effect is less prominent in three dimensions than in two or one. Unlike in the one-dimensional model, the force no longer varies monotonically with distance, but shows oscillations, reflecting the greater geometrical complexity of the lattice in the higher dimensions.
At low temperatures, the solubility of the solute is found to decrease with increasing temperature. This is another signature of the hydrophobic effect, and also agrees with what had been found in the one-dimensional model.
## Acknowledgements
BW is grateful to the Institute of Theoretical Physics and to the Debye Institute of the University of Utrecht for their hospitality in the fall of 1999. His work at Cornell was supported by the U.S. National Science Foundation and the Cornell Center for Materials Research. The authors thank M.D. Zeidler for calling their attention to the work in Reference 2. |
warning/0003/quant-ph0003107.html | ar5iv | text | # 1 Introduction
## 1 Introduction
In this paper we give a new proof, using only finite sums and avoiding the need for analytic methods, of the Landsberg-Schaar formula for quadratic Gauss sums. The key idea, which springs from earlier work of one of us , is to apply the quantization of a quantum mechanical system whose phase space is a torus, with both position and momentum (and also time) discrete and cyclic. On adapting Feynmanโs โsum over pathsโ method to this system a discrete and finite version of Kelvinโs method of images is obtained.
The formula states that for positive integers $`p`$ and $`q`$
$$\frac{1}{\sqrt{p}}\underset{n=0}{\overset{p1}{}}\mathrm{exp}\left(\frac{2\pi in^2q}{p}\right)=\frac{e^{\pi i/4}}{\sqrt{(2q)}}\underset{n=0}{\overset{2q1}{}}\mathrm{exp}\left(\frac{\pi in^2p}{2q}\right).$$
(1)
In number theory this formula plays a central rรดle, underpinning key results relating to quadratic reciprocity and characters. The formula also promises to be a useful adjunct to discrete Fourier transform methods used, for instance, in some algorithms for quantum computing . In this context it essentially provides the discrete inverse quantum Legendre transformation from the free Hamiltonian to the free Lagrangian.
The standard proof of the Landsberg-Schaar formula (as given, for example, in ) is obtained by putting $`\tau =2iq/p+ฯต`$, $`ฯต>0`$, in the Jacobi identity for the theta function:
$$\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\pi n^2\tau }=\frac{1}{\sqrt{\tau }}\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\pi n^2/\tau },$$
(2)
and then letting $`ฯต0`$. This method is an example of Heckeโs observation that โexact knowledge of the behaviour of an analytic function in the neighbourhood of its singular points is a source of arithmetic theoremsโ p225.
While this method certainly establishes the truth of the formula, Heckeโs insight notwithstanding it is somewhat unsatisfactory to have to use analytic methods and take limits when only finite sums are involved. The proof given in this paper does remain entirely in the arena of finite sums; it demonstrates the extraordinary range of Feynmanโs method of โsumming over historiesโ, which here extends to the discrete and cyclic regime in a pleasing and effective way. So, in a sense, the quantum mechanics offers another example of Heckeโs insight
The key ingredient in this paper is quantization on a toroidal phase space. A first step was taken by one of us in , leading to a proof of the Jacobi identity by using random walks to quantize a system with cylindrical phase space. The key idea here is the observation that the series $`\vartheta (it)=_{n=\mathrm{}}^+\mathrm{}e^{\pi n^2it}`$ is the trace of a certain quantum evolution operator $`\mathrm{exp}iHt/\mathrm{}`$, which converges in the distribution sense. Using path integral methods adapted to the cylinder, in particular Kelvinโs method of images , an alternative formula for the trace is found, and equating the two expressions gives the Jacobi identity. This work is described in section 2 as an introduction to the main result in this paper.
With a minor change of emphasis, the formula (1) turns up by taking the operator that gives (2) (for $`\tau =it`$) and then โlooking at the system at discrete times $`\tau =2i/p`$โ. This idea is pursued in section 3. In order to remove difficulties over convergence, and to obtain the Landsberg-Schaar formula (1), we change the problem to one in which instead of the infinite set of energy levels (7) there is only a finite number of energy levels $`E_n`$. This is achieved, following a suggestion of Berry , by replacing the cylindrical phase space (with angular position $`\theta `$ periodic but angular momentum $`L`$ unbounded) by a torus on which both $`\theta `$ and $`L`$ are periodic. On quantization both $`\theta `$ and $`L`$ are discrete, and moreover, as we see, discrete times are also required. Adapting the methods used in section 2 to prove the Jacobi identity to this new setting, again two expressions for the trace of the quantum evolution operator are obtained, one by working in the momentum basis in which the evolution operator is diagonal and the other by path integral methods (using a novel discrete variant on Kelvinโs method of images); the equality of these two expressions gives the Landsberg-Schaar identity.
This first description of the discrete and cyclic path integral method is presented somewhat heuristically, but in a manner which should emphasise the ideas and motivation. In section 4 full details of the toroidal quantization are given together with a justification of the normalisations used in section 3.
## 2 Cylindrical phase space and the Jacobi identity
In this section earlier work of one of us is described as a prelude to the main result of the paper, in order to demonstrate some of the novel features of the work.
Consider a rigid body constrained to rotate about a fixed axis (cf Schulman ). Let $`I`$ denote the moment of inertia and $`\theta `$ denote the angle of rotation (given in radians so that it is taken mod $`2\pi `$). An angular Schrรถdinger picture is used, with the angular momentum observable $`L`$ represented as the operator
$$L=i\mathrm{}\frac{}{\theta }$$
(3)
(where $`\mathrm{}`$ as usual stands for Planckโs constant $`h`$ divided by $`2\pi `$). Using the classical expression for the energy or Hamiltonian
$$H=\frac{L^2}{2I},$$
(4)
the Schrรถdinger equation for the wave function $`\psi (\theta ,t)`$ is then
$$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2I}\frac{^2\psi }{\theta ^2}.$$
(5)
Suppose that $`E_n`$ denotes the $`n^{th}`$ eigenvalue of the quantized Hamiltonian with energy eigenfunction $`\varphi _n`$,
$$\frac{\mathrm{}^2}{2I}\frac{^2}{\theta ^2}\varphi _n(\theta )=E_n\varphi _n(\theta ),$$
(6)
with respect to boundary values determined by periodicity in $`\theta `$ with period $`2\pi `$. Evidently
$$E_n=\frac{n^2\mathrm{}^2}{2I},\varphi _n(\theta )=e^{in\theta }.$$
(7)
The solution of (5) satisfying the initial condition
$$\psi (\theta ,0)=\dot{\delta }=\frac{1}{2\pi }\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{in(\theta \theta _0)},$$
(8)
where $`\dot{\delta }`$ denotes the periodic delta distribution with period $`2\pi `$ gives the kernel (or matrix element) of the evolution operator $`\mathrm{exp}iHt/\mathrm{}`$:
$$K(\theta ,t;\theta _0,0)=\theta |e^{\frac{iHt}{\mathrm{}}}|\theta _0=\frac{1}{2\pi }\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\frac{i\mathrm{}n^2t}{2I}}e^{in(\theta \theta _0)},$$
(9)
which is convergent in the distribution sense; alternatively one can give $`t`$ a small imaginary part to restore convergence. Later we shall see that when both position and momentum are cyclic, with phase space a torus, the problem of convergence does not arise.
From(9) we see that the trace of the evolution operator is
$$\mathrm{Tr}\left(\mathrm{exp}iHt/\mathrm{}\right)=\frac{1}{2\pi }\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\frac{i\mathrm{}n^2t}{2I}}$$
(10)
with convergence in the sense described above.
Now, as in , we follow adapted to the cylindrical phase space. The purpose of the section is to express the old idea of using the method of images and the universal cover of the circle in terms of path integrals, in such a way that the argument can be adapted to the toroidal phase space of section 3, obtaining a second expression for the trace of the evolution operator $`\mathrm{exp}iHt/\mathrm{}`$ which on comparison with (10) gives the Jacobi identity.
Denote by
$$S(\theta (t))=_0^t(\dot{\theta },\theta ;t)๐t,$$
(11)
the classical action, where $``$ is the Lagrangian, and then we obtain the evolution amplitude from $`(\theta _0,0)`$ to $`(\theta ,t)`$ as
$$K(\theta ,t;\theta _0,0)\underset{\text{periodic paths from }(\theta _0,0)\text{ to }(\theta ,t)}{}e^{iS(\theta )/hbar}$$
(12)
where โ$``$โ indicates that a normalisation factor (c.f. (13)) is required. In order to carry out the path integral sum we carry out the integration on the universal covering space, $``$, of $`S^1`$, that is, we lift the homotopy classes to the universal covering space . Corresponding to paths from $`\theta _0`$ to $`\theta `$ in $`S^1`$, we have paths from some fixed $`\theta _0^{}`$ in $`\nu ^1(\theta _0)`$ (here $`\nu `$ denotes the covering projection from $``$ to $`S^1`$) to each of the $`\theta _j^{}`$ in $`\nu ^1(\theta )`$, where the index $`j`$ runs through the fundamental group ($``$) of $`S^1`$. It follows (this is of course Kelvinโs method of images) that
$$K(\theta ,t;\theta _0,0)=\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}\left(\frac{I}{2\pi i\mathrm{}t}\right)^{\frac{1}{2}}\mathrm{exp}\left(\frac{iI}{2\mathrm{}t}(\theta \theta _02\pi n)^2\right).$$
(13)
(The normalisation factor is taken from .)
On equating the expressions for the trace in (10) and (13) we obtain
$$\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\frac{(i\mathrm{}n^2t)}{2I}}=\left(\frac{2\pi I}{i\mathrm{}t}\right)^{\frac{1}{2}}\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{\frac{2\mathrm{}^2n^2I}{\mathrm{}it}}.$$
(14)
(As already observed, care is needed over convergence, which is intended in the distributional sense.) If we arrange that $`2\pi I=\mathrm{}`$, $`\tau =it`$, then we obtain formally the usual Jacobi identity (2) for $`\vartheta (it)`$.
After this work was presented it was suggested by Berry that the way forward to obtaining the Landsberg-Schaar formula by similar means might be to consider a system where the angular momentum, as well as the angle, was cyclic so that the phase space was compact and the quantized observables would not only be discrete but also have a finite range. This approach, which does indeed lead to the Landsberg-Schaar formula, is developed in the following section.
## 3 Toroidal phase space and the Landsberg-Schaar formula
In this section we adapt the methods of the preceding section to a toroidal phase space on which both $`L`$ and $`\theta `$ are periodic. On quantization this gives discreteness to both these variables, and makes it possible (provided that suitable values are used for the various constants involved) to have only a finite number of energy eigenstates. Time, too, becomes quantized as emerges when evolution is considered. In this section we concentrate on ideas, obtaining the Landsberg-Schaar formula by proceeding in the manner of a physicist, while in the next section the quantization scheme is fully described and normalisation factors are derived.
Suppose that the phase space of our system is a torus, with angular momentum $`L`$ of period $`P`$ and angle of rotation $`\theta `$ of period $`2\pi `$. Quantization in $`L`$ and $`\theta `$, when both $`\theta `$ and $`L`$ are periodic, gives, respectively:
$$L=n\mathrm{},\theta =\frac{2\pi m\mathrm{}}{P},m,n.$$
(15)
The phase space is quantized to a lattice on the torus, and the small rectangles with sides $`\mathrm{}`$ and $`\frac{2\pi \mathrm{}}{P}`$ fit a whole number of times into the phase space torus area. The area of the phase space torus is $`2\pi P`$. So $`2\pi P=2\pi N\mathrm{}`$, ($`P=N\mathrm{},N`$) and so the small rectangles fit $`N^2`$ times into the phase space torus.
There are precisely $`N`$ energies, $`E_n,1nN`$, and so the trace of the evolution operator is given by:
$`\mathrm{Tr}(\mathrm{exp}iHt/\mathrm{})`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{n=1}{\overset{N}{}}}e^{\frac{iE_nt}{\mathrm{}}}`$ (16)
$`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{n=1}{\overset{N}{}}}e^{\frac{i\mathrm{}n^2t}{2I}}.`$
Now we turn to the $`L`$-periodicity and show that it implies a restriction on the times, $`t`$, of โlooking at the systemโ.
Any quantum mechanical state $`\psi (\theta ,t)`$ has an expansion:
$$\psi (\theta ,t)=\underset{k=\mathrm{}}{\overset{+\mathrm{}}{}}a_ke^{i\left(k\theta \frac{\mathrm{}k^2t}{2I}\right)},a_k=a_{k+N}$$
(17)
with the usual observations about convergence. Now take the Fourier transform in $`\theta `$ of (17) to go from position (angle) space to momentum space. We obtain (here $`\widehat{\psi }`$ denotes the Fourier transform in the variable $`\theta `$),
$$\widehat{\psi }(L,t)=_{\mathrm{}}^+\mathrm{}\psi (\theta ,t)e^{ip\theta /\mathrm{}}๐\theta ,$$
(18)
and we use $`\left[e^{ik\theta }\right]=\delta (Lk\mathrm{})`$)
$$\widehat{\psi }(L,t)=\underset{k=\mathrm{}}{\overset{+\mathrm{}}{}}a_ke^{\frac{i\mathrm{}k^2t}{2I}}\delta (Lk\mathrm{}),$$
(19)
where $`\delta `$ denotes the Dirac delta distribution, and
$`\widehat{\psi }(L,t)`$ $`=`$ $`{\displaystyle \underset{k=\mathrm{}}{\overset{+\mathrm{}}{}}}a_ke^{\frac{i\mathrm{}k^2t}{2I}}\delta (L+Pk\mathrm{})`$ (20)
$`=`$ $`{\displaystyle \underset{k=\mathrm{}}{\overset{+\mathrm{}}{}}}a_ke^{\frac{i\mathrm{}(k+N)^2t}{2I}}\delta (Lk\mathrm{}).`$
Thus periodicity of period $`P`$ in $`L`$, that is $`\widehat{\psi }(L+P,t)=\widehat{\psi }(L,t)`$, implies
$$1=e^{i\mathrm{}t(N^2+2kN)/(2I)}$$
(21)
so that (assuming $`N`$ is even, as is required later)
$$t=\frac{2\pi mI}{N\mathrm{}},m.$$
(22)
On substituting from (22) in (16) we obtain
$`\mathrm{Tr}(\mathrm{exp}iHt/\mathrm{})`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{n=0}{\overset{N}{}}}e^{i\mathrm{}n^2\frac{2\pi mI}{N\mathrm{}2I}}`$ (23)
$`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{n=0}{\overset{N}{}}}e^{i\mathrm{}\pi n^2\frac{m}{N}}.`$
We have to compute
$$\underset{\text{periodic paths}}{}e^{iS(\theta )/\mathrm{}},$$
(24)
where the sum is now over all periodic paths on the torus. At the $`n^{th}`$ energy level, $`p_n`$ is constant:
$$E_n=\frac{p_n^2}{2I},I\dot{\theta }_n=p_n,I\theta =p_nt.$$
(25)
Hence, for the โtime of lookingโ given in (22),
$$|I\theta |p_n.\frac{2|m|\pi I}{N\mathrm{}}|m|.2\pi I.$$
(26)
On repeating the argument given in section 2, but now restricting $`\theta `$ as in (26), we obtain in place of (13)
$$K(\theta ,t)=\underset{n=0}{\overset{m1}{}}\left(\frac{I}{2\pi i\mathrm{}t}\right)^{\frac{1}{2}}\mathrm{exp}\left(\frac{iI}{2\mathrm{}t}(\theta 2\pi n)^2\right),$$
(27)
where as before
$$t=\frac{2\pi mI}{N\mathrm{}}.$$
(28)
(The normalisation factor used is taken from (13), and is justified in section 4; while it might seem simpler to derive this by the method used by Davison than by the methods of section 4, in fact there is a crucial step (equation (3.5) of Davisonโs paper) which would require knowledge of the Gauss sums under study.)
It follows from (27) that the propagator trace is
$`\mathrm{Tr}(\mathrm{exp}iHt/\mathrm{})`$ $`=`$ $`\left({\displaystyle \frac{N}{4\pi ^2im}}\right)^{\frac{1}{2}}{\displaystyle \underset{n=0}{\overset{m1}{}}}\mathrm{exp}\left({\displaystyle \frac{in^2}{\left(\frac{2\mathrm{}.2m\pi }{N\mathrm{}}\right)}}\right)`$ (29)
$`=`$ $`\left({\displaystyle \frac{N}{im}}\right)^{\frac{1}{2}}{\displaystyle \underset{n=0}{\overset{m1}{}}}e^{i\pi n^2N/m}.`$
So, on equating (16) and (29), we obtain
$$\underset{n=0}{\overset{N}{}}e^{\frac{2\pi in^2}{N}}=\left(\frac{N}{im}\right)^{\frac{1}{2}}\underset{n=0}{\overset{m1}{}}e^{i\pi n^2N/m}.$$
(30)
Finally in (30) we take $`N=2q`$, $`m=p`$, $`i^1=e^{\pi i/2}`$ to obtain
$$\frac{e^{\pi i/4}}{\sqrt{(2q)}}\underset{n=0}{\overset{2q1}{}}\mathrm{exp}\left(\frac{\pi in^2p}{2q}\right)=\frac{1}{\sqrt{p}}\underset{n=0}{\overset{p1}{}}\mathrm{exp}\left(\frac{2\pi in^2q}{p}\right),$$
(31)
which is the Landsberg-Schaar formula (1).
## 4 Normalisation of the discrete path integral
In this section details are given of the quantum mechanical system whose path integration is used in the previous section to prove the Landsberg-Schaar formula (1). The necessary path integral formula is derived by adapting the original approach of Feynman to the discrete and cyclic setting. Because all sums are finite in this case there are no convergence or other analytic difficulties, so that a precise result is obtained with well-defined normalisation. This confirms the validity of the more heuristic, but geometrically and arithmetically well-motivated, use of the Feynman principle in the previous section.
The starting point is the classical phase space of the system, which is taken to be toroidal with coordinates $`\theta ,L`$ of periodicity $`2\pi `$ and $`P`$ respectively. As Hilbert space for our system we take $`=^N`$ (where $`N`$ is a positive integer) realised as
$$\{f:|f(\theta )=\underset{n=1}{\overset{N}{}}a_n\mathrm{exp}\left(\frac{2\pi in\theta }{N}\right)\}$$
where each $`a_n`$ is a complex number.
One basis of $``$ is then plainly $`\{f_k:k=0,\mathrm{},N1\}`$ with $`f_k(\theta )=\frac{1}{\sqrt{N}}\mathrm{exp}(\frac{2\pi ik\theta }{N})`$. This is the angular momentum basis, that is, each $`f_k`$ is an eigenvector of the angular momentum operator $`L`$ (defined now by $`L=i\frac{N}{2\pi }\frac{}{\theta }`$) with eigenvalue $`k`$. Using Dirac notation, we write $`f_k`$ as $`|k`$. For simplicity we use units in which $`\mathrm{}`$ takes the value $`1`$.
Another basis is $`\{b_r:r=1,\mathrm{},N\}`$ with
$$b_r(\theta )=\frac{1}{N}\underset{k=0}{\overset{N1}{}}\mathrm{exp}\left(\frac{2\pi i(\theta r)k}{N}\right).$$
(32)
The inner product on $``$ is defined by
$$f|g=\underset{j=0}{\overset{N1}{}}f(j){}_{}{}^{}g(j).$$
(33)
With this inner product both of these bases are orthonormal.
The $`b_r`$ may be regarded as the position basis if one restricts the domain of the elements $`f`$ of $``$ to $`_N=\{0,1,\mathrm{},N1\}`$ and defines the (exponentiated angular) position operator $`\widehat{x}`$ by
$$\widehat{x}f(\theta )=\mathrm{exp}\left(2\pi i\frac{\theta }{N}\right)f(\theta ),$$
(34)
since then
$$\widehat{x}b_r=\mathrm{exp}\left(2\pi i\frac{r}{N}\right)b_r$$
(35)
so that $`b_r`$ is an eigenstate of $`\widehat{x}`$, with eigenvalue $`\mathrm{exp}\left(2\pi i\frac{r}{N}\right)`$. (In Dirac notation we write $`b_r`$ as $`|r`$.)
Now the proof of the Landsberg-Schaar identity essentially involves calculating the trace of $`\mathrm{exp}iHt`$ (with $`H=\frac{L^2}{2I}`$ and $`t=\frac{2\pi mI}{N}`$ as before) in these two different bases. From the outset we will set $`m=p`$, where $`p`$ is one of the two integers in the Landsberg-Schaar formula.
Method 1 is the direct way, that is, working in the momentum basis in which $`H`$ is diagonal.
We have $`Ht|k=\frac{\pi k^2p}{N}|k`$ so that
$$\mathrm{Tr}\left(\mathrm{exp}iHt\right)=\underset{k=0}{\overset{N1}{}}\mathrm{exp}\left(\frac{i\pi k^2p}{N}\right)$$
(36)
Method 2 uses discrete, cyclic โpath integralsโ. We consider $`r|\mathrm{exp}iHt|s`$.
Breaking the time interval into $`p`$ steps $`\mathrm{\Delta }t=\frac{t}{p}`$, and using $`s_i`$ to label basis elements $`b_{s_i}`$ at the $`i^{th}`$ step, we have
$`r|\mathrm{exp}iHt|s`$ $`=`$ $`r|\mathrm{exp}iH\mathrm{\Delta }t|s_{p1}s_{p1}|\mathrm{exp}iH\mathrm{\Delta }t|s_{p2}`$ (37)
$`\mathrm{}s_1|\mathrm{exp}iH\mathrm{\Delta }t|s`$
(with summation from $`0`$ to $`N1`$ over each of the intermediate $`s_i,i=1,\mathrm{},p1`$). We thus need to consider $`s_i|\mathrm{exp}iH\mathrm{\Delta }t|s_{i1}`$. Now
$`s_i|\mathrm{exp}iH\mathrm{\Delta }t|s_{i1}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}s_i|kk|s_{i1}\mathrm{exp}\left({\displaystyle \frac{\pi ik^2}{N}}\right)`$ (38)
$`=`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}{\displaystyle \frac{1}{N}}\mathrm{exp}\left({\displaystyle \frac{2\pi ik(s_is_{i1})}{N}}\right)\mathrm{exp}\left({\displaystyle \frac{\pi ik^2}{N}}\right)`$
(using the fact that
$$r|k=\frac{1}{\sqrt{N}}\mathrm{exp}\frac{2\pi kr}{N}).$$
(39)
Thus
$`s_i|\mathrm{exp}iH\mathrm{\Delta }t|s_{i1}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}{\displaystyle \frac{1}{N}}\mathrm{exp}\left({\displaystyle \frac{\pi i}{N}}\left(k(s_is_{i1})\right)^2\right)`$ (40)
$`\mathrm{exp}\left(\pi i{\displaystyle \frac{(s_is_{i1})^2}{N}}\right).`$
Again, as in section 3, we set $`N=2q`$, so that in particular $`N`$ is even and we have
$`s_i|\mathrm{exp}iHT|s_{i1}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{N1}{}}}{\displaystyle \frac{1}{N}}\mathrm{exp}\left({\displaystyle \frac{2\pi i}{2N}}\left(k(s_is_{i1})\right)^2\right)\mathrm{exp}\left(2\pi i{\displaystyle \frac{(s_is_{i1})^2}{2N}}\right)`$ (41)
$`=`$ $`{\displaystyle \frac{1}{\sqrt{iN}}}\mathrm{exp}\left(2\pi i{\displaystyle \frac{(s_is_{i1})^2}{2N}}\right)`$
by equation (50) of the appendix. Thus
$$r|\mathrm{exp}iHT|s=(iN)^{p/2}\underset{s_1=0}{\overset{N1}{}}\mathrm{}\underset{s_{p1}=0}{\overset{N1}{}}\mathrm{exp}\left(2\pi i\frac{_{i=1}^p(s_is_{i1})^2}{2N}\right)$$
(42)
where $`s_0=r,s_p=s`$.
At this stage we assume that $`p`$ and $`q`$ are coprime. Then, if $`p`$ is odd, so that $`p`$ and $`N=2q`$ are coprime, we may observe that
$$\underset{l=0}{\overset{N1}{}}\underset{k=0}{\overset{p1}{}}\mathrm{exp}\left(2\pi i\frac{(kN+lp)^2}{2Np}\right)=\underset{t=0}{\overset{Np1}{}}\mathrm{exp}\frac{2\pi it^2}{2Np}=\sqrt{iNp}$$
(43)
where we have used the result (49) from the appendix. (Because $`p`$ and $`N`$ are coprime the expression $`kN+lp`$ takes all $`Np`$ distinct values ($`\mathrm{mod}Np`$) as $`k`$ ranges from $`0`$ to $`p1`$ and $`l`$ ranges from $`0`$ to $`N1`$.) Also, if $`p=2r`$ is even (but $`p`$ and $`q`$ are still coprime) then
$$\underset{l=0}{\overset{N1}{}}\underset{k=0}{\overset{p1}{}}\mathrm{exp}\left(2\pi i\frac{(kN+lp)^2}{2Np}\right)=\underset{l=0}{\overset{2q1}{}}\underset{k=0}{\overset{2r1}{}}\mathrm{exp}\left(2\pi i\frac{(kq+lr)^2}{2qr}\right)=\sqrt{(1+(1)^r)iNp}.$$
(44)
This result allows us to insert in the expression for $`r|\mathrm{exp}iHt|s`$ the extra summations $`_{l=0}^{N1}_{k=0}^{p1}\mathrm{exp}\left(2\pi i\frac{(kN+lp)^2}{2Np}\right)`$ together with the compensating factor $`1/\sqrt{iNp}`$ (if $`p`$ is odd) or $`1/\sqrt{2iNp}`$ (if $`p=2r`$ with $`r=2r^{}`$ even). We can overlook the case where $`p=2r`$ and $`r`$ is odd, since in that case the Landsberg-Schaar equation is trivially satisfied. This step effectively allows the winding number $`k`$ for a path to be shared over the $`p`$ steps in the path.
Using these facts, we see that if $`p`$ and $`q`$ are coprime and $`a=1`$ when $`p`$ is odd while $`a=2`$ when $`p=4r^{}`$,
$`r|\mathrm{exp}iHt|s`$
$`=`$ $`{\displaystyle \frac{(iN)^{p/2}}{\sqrt{iaNp}}}{\displaystyle \underset{s_1=0}{\overset{N1}{}}}\mathrm{}{\displaystyle \underset{s_{p1}=0}{\overset{N1}{}}}{\displaystyle \underset{l=0}{\overset{N1}{}}}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i\left({\displaystyle \underset{i=1}{\overset{p}{}}}{\displaystyle \frac{(s_is_{i1})^2}{2N}}+{\displaystyle \frac{(kN+lp)^2}{2Np}}\right)\right)`$
$`=`$ $`{\displaystyle \frac{(iN)^{p/2}}{\sqrt{iaNp}}}{\displaystyle \underset{s_1=0}{\overset{N1}{}}}\mathrm{}{\displaystyle \underset{s_{p1}=0}{\overset{N1}{}}}{\displaystyle \underset{l=0}{\overset{N1}{}}}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i\left({\displaystyle \underset{i=1}{\overset{p}{}}}\left({\displaystyle \frac{(s_is_{i1})^2}{2N}}+{\displaystyle \frac{l^2}{2N}}\right)+{\displaystyle \frac{(k^2q)}{p}}\right)\right)`$
$`=`$ $`{\displaystyle \frac{(iN)^{p/2}}{\sqrt{iaNp}}}{\displaystyle \underset{s_1=0}{\overset{N1}{}}}\mathrm{}{\displaystyle \underset{s_{p1}=0}{\overset{N1}{}}}{\displaystyle \underset{l=0}{\overset{N1}{}}}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i\left({\displaystyle \underset{i=1}{\overset{p}{}}}{\displaystyle \frac{\left(s_is_{i1}+l\right)^2}{2N}}+{\displaystyle \frac{(k^2q)}{p}}\right)\right)`$
provided that $`rs0(\mathrm{mod}N)`$.
Now let $`u_1=s_1s_0+l=s_1r+l`$, $`u_2=s_2s_1+l`$ and so on, with $`u_p=s_ps_{p1}+l=ss_{p1}+l`$. Then $`_{i=1}^pu_i=rs+lp`$ so that $`u_p=rs+lp_{i=1}^{p1}i_i`$. Thus if $`p`$ and $`N`$ are coprime and $`s_1,\mathrm{},s_{p1}`$ are fixed the integer variable $`u_p`$ will take each value $`0,1,\mathrm{},N1\mathrm{mod}N`$ precisely once as $`l`$ ranges from $`0`$ to $`N1`$, so that
$`r|\mathrm{exp}iHT|s`$ (46)
$`=`$ $`{\displaystyle \frac{(iN)^{p/2}}{\sqrt{iNp}}}{\displaystyle \underset{u_1=0}{\overset{N1}{}}}\mathrm{}{\displaystyle \underset{u_p=0}{\overset{N1}{}}}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i\left({\displaystyle \underset{i=1}{\overset{p}{}}}{\displaystyle \frac{u_i^2}{2N}}\right)\right)\mathrm{exp}\left(2\pi i{\displaystyle \frac{(k^2q)}{p}}\right)`$
$`=`$ $`{\displaystyle \frac{(iN)^{p/2}}{\sqrt{iNp}}}(Ni)^{p/2}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i{\displaystyle \frac{(k^2q)}{p}}\right)\text{by (}\text{49}\text{)}`$
$`=`$ $`{\displaystyle \frac{1}{\sqrt{ipN}}}{\displaystyle \underset{k=0}{\overset{p1}{}}}\mathrm{exp}\left(2\pi i{\displaystyle \frac{(k^2q)}{p}}\right),`$
Hence
$`\mathrm{Tr}\left(\mathrm{exp}iHt\right)`$ $`=`$ $`{\displaystyle \underset{s=0}{\overset{N1}{}}}s|\mathrm{exp}iHt|s`$ (47)
$`=`$ $`{\displaystyle \frac{N}{\sqrt{ipN}}}{\displaystyle \underset{k=1}{\overset{p}{}}}\mathrm{exp}\left(2\pi i{\displaystyle \frac{(k^2q)}{p}}\right)`$
$`=`$ $`{\displaystyle \frac{\sqrt{2q}}{\sqrt{ip}}}{\displaystyle \underset{k=1}{\overset{p}{}}}\mathrm{exp}\left(2\pi i{\displaystyle \frac{(k^2q)}{p}}\right)(\text{since }N=2q).`$
(If $`p=4r^{}`$ with $`p`$ and $`q`$ still coprime, then the values of $`u_p`$ are restricted, so that summation over $`l`$ contributes a factor $`\mathrm{exp}\left(\frac{u_i^2}{2N}\right)=\sqrt{2iNp}`$. The result above is thus also obtained in this case.)
Hence in all cases when $`p`$ and $`q`$ are coprime we have (using (36),
$$\underset{k=0}{\overset{2q1}{}}\mathrm{exp}\left(\pi i\frac{k^2p}{2q}\right)=\frac{\sqrt{2q}}{\sqrt{ip}}\underset{k=0}{\overset{p1}{}}\mathrm{exp}\left(2\pi i\frac{(k^2q)}{p}\right)$$
(48)
which immediately gives the Landsberg-Schaar formula when $`p`$ and $`q`$ are coprime. The general case follows on setting $`p=mp^{}`$, $`q=mq^{}`$ with $`p^{},q^{}`$ coprime. Appendix The two formulae below may be proved by elementary means:
$$\underset{n=0}{\overset{2r1}{}}\mathrm{exp}\left(\frac{2\pi i(ns)^2}{4r}\right)=\sqrt{2ri}$$
(49)
$$\underset{k=0}{\overset{2r1}{}}\mathrm{exp}\left(2\pi i\frac{(ks)^2}{4r}\right)=\sqrt{\frac{2r}{i}}$$
(50)
Given such an elementary evaluation of the Gauss sum, together with an elementary proof of the quadratic reciprocity law
$$\left(\frac{p}{q}\right)\left(\frac{p}{q}\right)=(1)^{\frac{p1}{2}\frac{q1}{2}},$$
(51)
where $`\left(\frac{p}{q}\right)`$ denotes the Legendre symbol and $`p`$ and $`q`$ are odd primes, one could prove (1) by evaluating each side and then appealing to the reciprocity law, and its extensions (to $`p=2`$ and $`p=1`$). Whether such a proof would provide insight into why the result is true is another matter.
Again, there is another proof of the theta function identity (2), due to Polya , which depends on an identity involving binomial coefficients, identities which are in turn related to the Markov chain approach to diffusion processes considered in .
The approach adopted in this paper presupposes an elementary proof of the evaluation of Gauss sums in (49) and (50), but not of the reciprocity law. It is perhaps tempting therefore to regard the present approach as a substitute, in some sense, to the reciprocity law, but we prefer to see it as a quantum mechanical equivalent of Heckeโs observation, in which Feynmanโs โsum over historiesโ in discrete time replaces the limiting process that derives (1) from (2). |
warning/0003/hep-th0003022.html | ar5iv | text | # Stable non-BPS D-branes of type I
## Abstract:
We review the boundary state description of the non-BPS D-branes in the type I string theory and show that the only stable configurations are the D-particle and the D-instanton. We also compute the gauge and gravitational interactions of the non-BPS D-particles and compare them with the interactions of the dual non-BPS particles of the heterotic string finding complete agreement. In this way we provide further dynamical evidence of the heterotic/type I duality.
D-branes, Boundary states conference: TMR meeting, Paris, 1999
Dirichlet branes (or D-branes for short) are a key ingredient in our understanding of the duality relations between superstring theories. They are described by a boundary conformal field theory, and admit a two-fold interpretation: on the one hand, the D-branes are objects on which open strings can end, and on the other hand they can emit or absorb closed strings . These two descriptions can be related to each other by world-sheet duality.
Introducing D-branes in a theory of closed strings amounts to extend their conformal field theory by introducing world-sheets with boundaries and imposing appropriate boundary conditions on the closed string coordinates $`X^\mu `$. In the operator formalism these boundary conditions are implemented through the so called boun-dary state <sup>1</sup><sup>1</sup>1For a recent review on the boundary state formalism and its applications, see Ref. $`|Dp`$, whose bosonic part is defined by the following eigenvalue problem
$`_\tau X^\alpha (\sigma ,0)|Dp_X`$ $`=`$ $`0,`$
$`\left(X^i(\sigma ,0)x^i\right)|Dp_X`$ $`=`$ $`0,`$ (1)
where the index $`\alpha =0,\mathrm{},p`$ labels the longitudinal directions, the index $`i=p+1,\mathrm{},9`$ labels the transverse directions and the $`x^i`$โs denote the position of the brane in the transverse space. World-sheet supersymmetry requires that analogous equations must be also imposed on the left and right moving fermionic fields $`\psi ^\mu `$ and $`\stackrel{~}{\psi }^\mu `$. These equations, which define the fermionic part of the boundary state, are
$`\left(\psi ^\alpha (\sigma ,0)\mathrm{i}\eta \stackrel{~}{\psi }^\alpha (\sigma ,0)\right)|Dp,\eta _\psi =0,`$
$`\left(\psi ^i(\sigma ,0)+\mathrm{i}\eta \stackrel{~}{\psi }^i(\sigma ,0)\right)|Dp,\eta _\psi =0,`$ (2)
where $`\eta =\pm 1`$. Notice that there are two consistent implementations of the fermionic boundary conditions corresponding to the sign of $`\eta `$, and consequently there are two different boundary states
$$|Dp,\eta =|Dp_X|Dp,\eta _\psi $$
(3)
both in the NS-NS and in the R-R sectors. The overlap equations (1) and (2) allow to determine the explicit structure of the boundary states (3) up to an overall factor. This normalization can then be uniquely fixed by factorizing amplitudes with closed strings emitted from a disk and turns out to be given by (one half of) the brane tension measured in units of the gravitational coupling constant, i.e.
$$T_p=\sqrt{\pi }\left(2\pi \sqrt{\alpha ^{}}\right)^{3p}.$$
We would like to remark that even if each boundary state $`|Dp,\eta `$ is perfectly consistent from the conformal field theory point of view, not all of them are acceptable in string theory. In fact, to describe a physical D-brane a boundary state has to satisfy three requirements :
* to be invariant under the closed string GSO projection (and also under orbifold or orientifold projections if needed);
* the tree level amplitude due to the exchange of closed strings between two boundary sta-tes, after modular transformation in the open string channel, has to make sense as a consistent open string partition function at one-loop;
* the open strings introduced through the D-branes must have consistent couplings with the original closed strings <sup>2</sup><sup>2</sup>2This last condition, which is rather difficult to prove in general, does not give more constraints than the first two in the case of type I and II theories..
Using these prescriptions, it is rather simple to find the boundary state for the supersymmetric BPS D$`p`$-branes of type II. In particular, the GSO projection of the type II theories forces us to retain only the following linear combinations
$`|Dp_{\mathrm{NS}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[|Dp,+_{\mathrm{NS}}|Dp,_{\mathrm{NS}}\right]`$
$`|Dp_\mathrm{R}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[|Dp,+_\mathrm{R}+|Dp,_\mathrm{R}\right]`$ (4)
in the NS-NS and in the R-R sectors respectively, with $`p=0,2,4,6,8`$ for IIA, $`p=1,1,3,5,7,9`$ for IIB. The normalization of the boundary states (4) can be deduced by requiring that the spectrum of the open strings living on the D$`p`$-brane (called $`p`$-$`p`$ strings) be supersymmetric. To read the spectrum of these open strings from the boundary state, one has first to evaluate the closed string exchange amplitude
$$Dp|P|Dp,$$
where $`P`$ is the closed string propagator
$$P=\frac{\alpha ^{}}{2}_0^{\mathrm{}}๐t\mathrm{e}^{t(L_0+\stackrel{~}{L}_02a)}$$
(with $`a_{\mathrm{NS}}=1/2`$ and $`a_\mathrm{R}=0`$), and then perform the modular transformation $`t1/s`$ to exhibit the open string channel. Applying this procedure, one finds the following relations
$`{}_{\mathrm{NS}}{}^{}Dp,\eta |P|Dp,\eta _{\mathrm{NS}}^{}`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}\mathrm{Tr}_{\mathrm{NS}}q^{2L_01},`$
$`{}_{\mathrm{NS}}{}^{}Dp,\eta |P|Dp,\eta _{\mathrm{NS}}^{}`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}\mathrm{Tr}_\mathrm{R}q^{2L_0},`$
$`{}_{\mathrm{R}}{}^{}Dp,\eta |P|Dp,\eta _{\mathrm{R}}^{}`$
$`={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}`$ $`\mathrm{Tr}_{\mathrm{NS}}(1)^Fq^{2L_01},`$ (5)
$`{}_{\mathrm{R}}{}^{}Dp,\eta |P|Dp,\eta _{\mathrm{R}}^{}`$
$`={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}`$ $`\mathrm{Tr}_\mathrm{R}(1)^Fq^{2L_0}=0,`$
where $`q=\mathrm{e}^{\pi s}`$. It is then clear that in order to obtain the supersymmetric (i.e. GSO projected) open string amplitude
$`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}[\mathrm{Tr}_{\mathrm{NS}}\left({\displaystyle \frac{1+(1)^F}{2}}\right)q^{2L_01}`$
$`\mathrm{Tr}_\mathrm{R}\left({\displaystyle \frac{1+(1)^F}{2}}\right)q^{2L_0}]`$ (6)
one must consider the following boundary state
$$|Dp=|Dp_{\mathrm{NS}}\pm |Dp_\mathrm{R}$$
(7)
where the sign ambiguity is related to the existence of branes and anti-branes. Note that both the NS-NS and the R-R components of the boundary state (7) have the same normalization so that the tension of a D$`p`$-brane essentially equals the density of its charge under the R-R potential: this is the BPS relation which is typical of the supersymmetric and stable branes of type II.
The criteria $`i)`$ \- $`iii)`$ defining physical D-bra-nes do not rely at all on space-time supersymmetry, and thus one may wonder whether in type II theories there may exist also non-supersymmetric branes. This problem has been systematically addressed in a series of papers by A. Sen \- , who constructed explicit examples of non-BPS (and hence non-supersymmetric) branes. In particular in Ref. , he considered the superposition of a D-string of type IIB and an anti-D-string (with a $`Z_2`$ Wilson line on it) and by suitably condensing the tachyons of the open strings stretching between the brane and the anti-brane, he managed to construct a new configuration of type IIB which behaves like a D-particle, does not couple to any R-R field and is heavier by a factor of $`\sqrt{2}`$ than the BPS D-particle of the IIA theory. The boundary state for this non-BPS D-brane has been explicitlely constructed in Ref. . This construction can be obviously generalized to the case of a pair formed by two BPS D$`(p+1)`$-branes with opposite R-R charge (and with a $`Z_2`$ Wilson line) which, after tachyon condensation, becomes a non-BPS D$`p`$-brane. Alternatively, this same non-BPS configuration can be described starting from a superposition of two BPS D$`p`$ branes with opposite R-R charge and modding out the theory by the operator $`(1)^{F_L}`$ whose effect is to change the sign of all states in the R-R and R-NS sectors. In this second scheme, a superposition of a D$`p`$-brane and anti-D$`p`$-brane of type IIA (IIB) becomes in the reduced theory a non-BPS D$`p`$-brane of type IIB (IIA). In either way we therefore find that there exist non-BPS D$`p`$-branes for $`p=0,2,4,6,8`$. For reviews on this subject, see Refs. .
These branes are manifestly non-supersym-metric, but nevertheless they satisfy the conditions $`i)`$ \- $`\mathrm{๐๐๐})`$ mentioned above, and thus are perfectly consistent from the closed string point of view. In particular, since they are not charged under any R-R field, the boundary state for these non-BPS D-branes has only the NS-NS component, namely
$$|Dp=\mu _p|Dp_{\mathrm{NS}}$$
(8)
where we have introduced a (positive) coefficient $`\mu _p`$ to allow for a different normalization with respect to the standard BPS case. This normalization can be deduced by requiring that the closed string amplitude between two non-BPS branes, after a modular transformation, has the interpretation of the partition function of a non-supersymmetric (i.e. without the GSO projection <sup>3</sup><sup>3</sup>3Note that the non-supersymmetric GSO projection $`(1(1)^F)/2`$ cannot correspond to a single brane since all NS open string zero-modes are projected out. ) open string model. Indeed, by requiring that
$$Dp|P|Dp=_0^{\mathrm{}}\frac{ds}{s}\left[\mathrm{Tr}_{\mathrm{NS}}q^{2L_01}\mathrm{Tr}_\mathrm{R}q^{2L_0}\right]$$
(9)
we find that $`\mu _p=\sqrt{2}`$, thus confirming that the non-BPS D-branes are heavier by a factor of $`\sqrt{2}`$ than the corresponding BPS ones. Although the-se non-BPS D-branes of type II may have interesting properties , it is clear from (9) that they are not stable, because the absence of the GSO projection on the open strings leaves the NS tachyon on their world-volume. However, these non-BPS branes could become stable in an orbifold of the type II theory, say IIA(B)$`/๐ซ`$, provided that the tachyon be odd under the projection $`๐ซ`$. In the orbifold theory, the non-BPS vacuum amplitude of the $`p`$-$`p`$ open-strings is clearly given by
$`๐ต_{open}=n{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}[\mathrm{Tr}_{\mathrm{NS}}\left({\displaystyle \frac{1+๐ซ}{2}}q^{2L_01}\right)`$
$`\mathrm{Tr}_\mathrm{R}\left({\displaystyle \frac{1+๐ซ}{2}}q^{2L_0}\right)]`$ (10)
where $`n`$ is a positive integer representing some possible multiplicity. (In our present discussion we take $`n=1`$ for the sake of simplicity, but the case $`n=2`$ will appear later.) The natural question to ask now is to which boundary state the amplitude (10) could correspond. In the case of a space-time orbifold, the perturbative spectrum of the bulk theory contains only closed strings which can be untwisted (U) or twisted (T) under the orbifold. Therefore, there are four sectors to which the bosonic states belong, namely (NS-NS;U), (R-R;U), (NS-NS;T) or (R-R;T), and the-re exist different types of boundary states depending on which components in those sectors they have. For example, when the orbifold projection $`๐ซ`$ acts as the inversion of some space-time coordinates, the boundary state which gives rise to (10) turns out to have only a component in the unstwisted NS-NS sector and another in the twisted R-R sector, i.e.
$$|Dp=\frac{1}{\sqrt{2}}\left(\sqrt{2}|Dp_{\mathrm{NS};\mathrm{U}}+\sqrt{2}|Dp_{\mathrm{R};\mathrm{T}}\right).$$
(11)
In particular, if one computes the exchange amplitude $`Dp|P|Dp`$ with this boundary state and performs a modular transformation, one can see that the terms of (10) with $`๐ซ`$ originate precisely from the twisted part of the boundary state.
In the case of a world-sheet orbifold, however, this simple picture does not hold. To illustrate this point, we consider the specific case of the type I theory which is the orbifold of the IIB theory by the world-sheet parity $`\mathrm{\Omega }`$ . The distinctive feature of this model is that the perturbative states of the twisted sector of the bulk theory now correspond to unoriented open strings which should then be appropriately incorporated in the boundary state formalism. Let us briefly summarize how this is done (for more details see ). The starting point is the projection of the closed string spectrum onto states which are invariant under $`\mathrm{\Omega }`$. The corresponding closed string partition function is obtained by adding a Klein bottle contribution to the modular invariant (halved) torus contribution. The Klein bottle is a genus one non-orientable self-intersecting surface which may be seen equivalently as a cylinder ending at two crosscaps. A crosscap is a line of non-orientability, a circle with opposite points identified, and thus the associated crosscap state $`|C`$ is defined by
$`X^\mu (\sigma +\pi ,0)|C`$ $`=`$ $`X^\mu (\sigma ,0)|C,`$ (12)
$`_\tau X^\mu (\sigma +\pi ,0)|C`$ $`=`$ $`_\tau X^\mu (\sigma ,0)|C,`$
and by the analogous relations appropriate for world-sheet fermions. As is clear from these equations, the crosscap state does not have any space-time interpretation but nevertheless it is related to the boundary state of the BPS space-time filling D9 brane through
$$|C\mathrm{i}^{L_0+\stackrel{~}{L}_0}|D9.$$
(13)
The normalization of $`|C`$, which may be fixed up to an overall sign using the action of $`\mathrm{\Omega }`$ on the massless closed string modes and the world-sheet duality, turns out to be 32 times the normalization of the boundary state for the D9-brane. Consequently, the (negative) charge for the unphysical 10-form R-R potential created by the crosscap must be compensated by the introduction of 32 D9 branes. In this way we then introduce unoriented open strings which start and end on these 32 D9 branes, whose vacuum amplitude is given by
$`๐ต_{open}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(2^{10}D9|P|D9`$
$`+\mathrm{\hspace{0.17em}\hspace{0.17em}2}^5D9|P|C+2^5C|P|D9),`$
where the first line represents the contribution of the annulus and the second line the contribution of the Mรถbius strip. By adding to (Stable non-BPS D-branes of type I) the contribution of the Klein bottle we obtain a modular invariant expression, in which the tadpoles for the massless unphysical states cancel if and only if we choose the still unfixed overall sign in front of the crosscap state to be +. A moment thought shows that this corresponds to choose the open string gauge group to be $`SO(32)`$ (the other sign instead leads to the gauge group $`Sp(32)`$). Thus, we can say that the type I theory possesses a โbackgroundโ boundary state given by
$$\frac{1}{\sqrt{2}}\left(|C+32|D9\right).$$
(15)
where the factor of $`1/\sqrt{2}`$ has been introduced to obtain the right normalization of the various spectra. Performing a modular transformation, we can rewrite the amplitude $`๐ต_{open}`$ of eq. (Stable non-BPS D-branes of type I) in the open string channel as follows
$`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{s}}[\mathrm{Tr}_{\mathrm{NS}}\left({\displaystyle \frac{1+(1)^F}{2}}{\displaystyle \frac{1+\mathrm{\Omega }}{2}}q^{2L_01}\right)`$
$`\mathrm{Tr}_\mathrm{R}\left({\displaystyle \frac{1+(1)^F}{2}}{\displaystyle \frac{1+\mathrm{\Omega }}{2}}q^{2L_0}\right)],`$
where the part depending on $`\mathrm{\Omega }`$ comes from the Mรถbius contribution. Thus, we see that in the type I theory the crosscap state plays the same role that the twisted part of the boundary state had in the space-time orbifolds. In the following, we shall use this remark in order to classify the stable non-BPS branes of type I theory.
We have seen before that the type IIB theory contains unstable non-BPS D$`p`$-branes with $`p=0,2,4,6,8`$ which are described by the boundary state (8). Now, we address the question whether these D-branes become stable in the type I theory, i.e. we examine whether the tachyons of the $`p`$-$`p`$ open strings are removed by $`\mathrm{\Omega }`$. As explained in , the world-sheet parity can be used to project the spectrum of the $`p`$-$`p`$ strings only if $`p=0,4,8`$. Thus, the non-BPS D2 and D6 branes will not be further considered. However, in order to be exhaustive, we must take into account also another kind of configuration, namely the superposition of a D$`p`$-brane and an anti-D$`p`$-brane of type IIB. This pair clearly does not carry any R-R charge, is represented by a boundary state of the form (8) and is unstable due to the presence of tachyons in the open strings stretching between the brane and the anti-brane. In the type I theory, however, these tachyons might be projected out. A systematic analysis shows that in this case $`\mathrm{\Omega }`$ can be used as a projection only if $`p=1,3,7`$.
In conclusion, we have to analyze the stability of the non-BPS D$`p`$-branes of type I with $`p=1,0,3,4,7,8`$ whose corresponding boundary states $`|Dp`$ are given by eq. (8) with suitable values of $`\mu _p`$. To address this problem, we need to consider the spectrum of the unoriented strings living on the brane world-volume (the $`p`$-$`p`$ sector), and also the spectrum of the open strings stretched between the D$`p`$-brane and each one of the 32 D9-branes of the background (the $`p`$-$`99`$-$`p`$ sector), in which tachyonic modes could develop.
Let us first analyze the $`p`$-$`p`$ sector, whose total vacuum amplitude is given by
$$๐_{\mathrm{tot}}=\frac{1}{2}\left(๐++^{}\right)$$
(16)
where $`๐`$ and $``$ are respectively the annulus and the Mรถbius strip contributions
$$๐=Dp|P|Dp\text{and}=Dp|P|C.$$
(17)
After a modular transformation, in the open string channel these amplitudes read respectively
$`๐`$ $`=`$ $`\mu _p^2V_{p+1}(8\pi ^2\alpha ^{})^{\frac{p+1}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{2s}}s^{\frac{p+1}{2}}`$ (18)
$`\times `$ $`\left[{\displaystyle \frac{f_3^8(q)f_2^8(q)}{f_1^8(q)}}\right],`$
and
$``$ $`=`$ $`2^{\frac{7p}{2}}\mu _pV_{p+1}(8\pi ^2\alpha ^{})^{\frac{p+1}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{2s}}s^{\frac{p+1}{2}}`$ (19)
$`[\mathrm{e}^{\mathrm{i}(p9)\pi /4}{\displaystyle \frac{f_4^{p1}(\mathrm{i}q)f_3^{9p}(\mathrm{i}q)}{f_1^{p1}(\mathrm{i}q)f_2^{9p}(\mathrm{i}q)}}`$
$`\mathrm{e}^{\mathrm{i}(9p)\pi /4}{\displaystyle \frac{f_3^{p1}(\mathrm{i}q)f_4^{9p}(\mathrm{i}q)}{f_1^{p1}(\mathrm{i}q)f_2^{9p}(\mathrm{i}q)}}],`$
where $`f_1`$, $`f_2`$ $`f_3`$ and $`f_4`$ are the standard one-loop functions defined for example in Ref. . The spectrum of the $`p`$-$`p`$ open strings can be analyzed by expanding the total amplitude $`๐_{\mathrm{tot}}`$ in powers of $`q`$. The leading term in this expansion is
$`๐_{\mathrm{tot}}`$ $``$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{2s}}s^{\frac{p+1}{2}}q^1`$ (20)
$`\times \left[\mu _p^22\mu _p\mathrm{sin}\left({\displaystyle \frac{\pi }{4}}(9p)\right)\right].`$
The $`q^1`$ behavior of the integrand signals the presence of tachyons in the spectrum; therefore, in order not to have them, we must require that
$$\mu _p=2\mathrm{sin}\left(\frac{\pi }{4}(9p)\right).$$
(21)
Since $`\mu _p`$ has to be positive, the only possible solutions are
$$\begin{array}{ccccc}& & & & \\ p& 1& 0& \mathrm{\hspace{0.17em}7}& 8\\ & & & & \\ \mu _p& 2& \sqrt{2}& 2& \sqrt{2}\end{array}$$
(22)
From this table we see that in the type I theory there exist two even non-BPS but stable D$`p`$-branes: the D-particle and the D8-brane. Both of them have a tension that is a factor of $`\sqrt{2}`$ bigger than the corresponding BPS branes of the type IIA theory. Moreover, there exist two odd non-BPS but stable D$`p`$-branes of type I: the D-instanton and the D7-brane. Their tension is twice the one of the corresponding type IIB branes, in accordance with the fact that, as mentioned above, they can be simply interpreted as the superposition of a brane with an anti brane, so that the R-R part of the boundary state cancels while the NS-NS part doubles.
This classification of the stable non-BPS D-branes of type I based on the table (22) is in complete agreement with the results of Refs. derived from the K-theory of space-time.
Let us now analyze the $`p`$-$`99`$-$`p`$ sector. The relevant quantity to consider is the โmixedโ cylinder amplitude
$$๐_{\mathrm{mix}}=\frac{32}{2}\left(Dp|P|D9+D9|P|Dp\right),$$
(23)
which, after a modular transformation in the open string channel, reads
$`๐_{\mathrm{mix}}`$ $`=`$ $`2^5\mu _pV_{p+1}(8\pi ^2\alpha ^{})^{\frac{p+1}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{ds}{2s}}s^{\frac{p+1}{2}}`$
$`\times `$ $`\left[{\displaystyle \frac{f_3^{p1}(q)f_2^{9p}(q)}{f_1^{p1}(q)f_4^{9p}(q)}}{\displaystyle \frac{f_2^{p1}(q)f_3^{9p}(q)}{f_1^{p1}(q)f_4^{9p}(q)}}\right],`$
where the first and second term in the square brackets account respectively for the NS and R sector. This expression needs some comments. First, for $`p=1,0`$ we see that there are no tachyons in the spectrum; moreover, the values of $`\mu _p`$ for the D-instanton and D-particle are crucial in order to obtain a sensible partition function for open strings stretching between the non-BPS objects and the 32 D9-branes. Indeed, they are the smallest ones that make integer the coefficients in the partition functions. Secondly, for $`p=7,8`$ we directly see the existence of a NS tachyon, so that the corresponding branes are actually unstable . Hence, only the D-instanton and the D-particle are fully stable configuration of type I string theory . Nevertheless, the strict relation connecting the D0-brane and the D(-1)-brane to the D8-brane and the D7-brane respectively suggests however that also the latter may have some non trivial meaning. Finally, we observe that the zero-modes of the Ramond sector of these $`p`$-$`99`$-$`p`$ strings are responsible for the degeneracy of the non-BPS D$`p`$-branes under the gauge group SO(32): in particular the D-particle has the degeneracy of the spinor representation of SO(32), as discussed in . Thus the D-particle accounts for the existence in type I of the non-perturbative non-BPS states required by the heterotic/type I duality.
These same methods may be used in order to study the stability of a non-BPS D$`p`$-brane in presence of another D$`q`$-brane. Indeed the spectrum of open strings stretching between two such (distant) objects at rest has a vacuum amplitude given by
$$\frac{\mu _p\mu _q}{2}({}_{\mathrm{NS}}{}^{}Dp|P|Dq_{\mathrm{NS}}^{}+{}_{\mathrm{NS}}{}^{}Dq|P|Dp_{\mathrm{NS}}^{}).$$
(24)
The overall factor of one-half indicates that, respectively to the IIB case, only the $`\mathrm{\Omega }`$ symmetric combinations are retained. By explicitly computing this amplitude, one can see that for $`|pq|3`$ and for sufficiently small values of the distance between the branes, a NS tachyon develops in the open string spectrum, thus signalling the unstability of the configuration. As a first consequence of this, we can conclude that the superposition of two non-BPS D-particles with trivial quantum numbers, decays into the vacuum . As a matter of fact, a stable non-BPS D$`p`$-brane is its own anti-brane, as may be also inferred from the K-theory analysis which shows that the conserved D-brane charge in that case is $`Z_2`$ valued . This analysis shows that there is no hope to form a stable superposition of $`N`$ non-BPS D$`p`$-branes of type I as they always exert an attractive force on each other, as may be seen from (24). This is to be contrasted with the case of a space-time orbifold where, for some particular values of the compactification radii, a compensation occurs between the attractive force due to exchange of untwisted NS-NS states and the repulsive force due to exchange of twisted R-R states. As a second consequence, and for analogous reasons, the superposition of a D-particle and of a D1-string is unstable and decays in the vacuum. Note that in that case, the 0-1 open string is T-dual of the 8-9 string responsible for the unstability of the D8-brane.
Up to now we have investigated the flat ten dimensional case. However, this analysis can be easily extended also to the case in which some directions are compactified. Contrarily to the BPS D-branes, the non-BPS one are not stable in all moduli space. As an example, let us consider the non-BPS D-particle and compactify one space direction along a circle of radius $`R`$. Then, one can observe that a tachyon develops in the $`0`$-$`0`$ open string sector if $`R<R_c=\sqrt{\alpha ^{}/2}`$, so that below the critical radius $`R_c`$ the configuration is unstable. The corresponding stable non-BPS configuration which carries the same quantum numbers in this range of moduli is a superposition of wrapped D1 and anti-D1 strings with a $`Z_2`$ Wilson line. Notice that when the time direction is compactified, the D-particle is stable for any value of the radius.
We now present the basic ideas and results about the gravitational and gauge interactions of two stable non-BPS D-particles of type I string theory (the detailed calculations and analysis of these interactions can be found in ).
In the type I theory, D-branes interact via exchanges of both closed and open bulk strings. Since the dominant diagram for open strings has the topology of a disk, it gives a subleading (in the string coupling constant) contribution to the diffusion amplitude of two branes which is thus dominated by the cylinder diagram, i.e. by the exchange of closed strings. In the long distance limit, this accounts for the gravitational interactions. Let us now use this observation to calculate the dominant part of the scattering amplitude between two D-particles of type I moving with a relative velocity $`v`$. This process can be simply analyzed using the boundary state formalism as explained in . What we need to compute is the cylinder amplitude between the boundary state of a static D-particle $`|D0`$ and the boundary state of a moving D-particle $`|D0,v`$. The latter is simply obtained by acting with a Lorentz boost on $`|D0`$, i.e.
$$|D0,v=\mathrm{e}^{\mathrm{i}\pi \theta J_{01}}|D0,$$
(25)
where $`\theta `$ is the rapidity along direction of motion (which we have taken to be $`x^1`$) defined through $`v=\mathrm{tanh}\pi \theta `$, and $`J_{01}`$ is the corresponding Lorentz generator. Thus, the amplitude we are looking for is
$$๐=D0|P|D0,v+D0,v|P|D0,$$
(26)
which indeed reduces to (24) for $`v0`$. From this expression, we can extract the long range gravitational potential energy, which, in the non relativistic limit, reads
$`V^{\mathrm{grav}}(r)=(2\kappa _{10})^2{\displaystyle \frac{M_0^2}{7\mathrm{\Omega }_8r^7}}`$
$`\times \left(1+{\displaystyle \frac{1}{2}}v^2+o(v^2)\right),`$ (27)
where $`r`$ is the radial coordinate, $`\mathrm{\Omega }_8`$ is the area of the unit $`8`$-dimensional sphere, $`M_0=T_0/\kappa _{10}`$ is the D-particle mass and $`\kappa _{10}`$ is the gravitational coupling constant in ten dimensions. Hence the boundary state calculation correctly reproduces the gravitational potential we expect for a pair of D-particles in relative motion.
Although they are subdominant in the string coupling constant, the interactions of the D-parti-cle with the open strings of the bulk are nevertheless interesting because they account for the gauge interactions. Since the non-BPS D-particles of type I are spinors of $`SO(32)`$, their gauge coupling is fixed by the spinorial representation they carry (except possibly by the overall strength). The stringy description of such a coupling has been provided in where we have shown that it is represented by an open string diagram with the topology of a disk with two boundary components, one lying on the D9-branes from which the gauge boson is emitted, and the other lying on the D-particle (see Figure 1).
At the points where the two boundary components join, we thus have to insert a vertex operator $`V_{90}`$ (or $`V_{09}`$) that induces the transition from Neumann to Dirichlet (or from Dirichlet to Neumann) boundary conditions in the nine space directions. As we have mentioned before, the $`SO(32)`$ degeneracy of the D-particle is due to the fermionic massless modes of the open strings stretching between the D-particle and each of the 32 D9-branes; therefore it is natural to think that the boundary changing operators $`V_{90}`$ and $`V_{09}`$ are given by the vertex operators for these massless fermionic modes . By construction, these operators carry Chan Paton factors in the fundamental representation of SO(32), while the vertex operator $`V_{\mathrm{gauge}}`$ for the gauge boson carries a Chan-Paton factor in the adjoint. As a consequence, the diagram represented in Figure 1 must be considered as the one point function of the gauge boson in the background formed by a D-particle seen as an object in the bi-fundamental representation of $`SO(32)`$. Hence, we do not see the entire gauge degeneracy of the D-particle because the degrees of freedom we use to describe it are not accurate enough. This is reminiscent from the fact that, in the boundary state formalism, also the Lorentz degeneracy of a D-brane is hidden. Using this result, we can easily compute the Coulomb potential energy $`V^{\mathrm{gauge}}(r)`$ for two D-particles placed at a distance $`r`$. Indeed, this is simply obtained by gluing two diagrams like that of Figure 1 with a gauge boson propagator, and its explicit expression turns out to be
$`V^{\mathrm{gauge}}(r)={\displaystyle \frac{g_{\mathrm{YM}}^2}{2}}{\displaystyle \frac{1}{7\mathrm{\Omega }_8r^7}}`$
$`\times \left(\delta ^{AB}\delta ^{CD}\delta ^{AC}\delta ^{DB}\right),`$ (28)
where $`g_{\mathrm{YM}}`$ is the gauge coupling constant in ten dimensional type I string theory, and $`A`$, $`B`$, $`C`$ and $`D`$ are indices in the fundamental representation of $`SO(32)`$.
We conclude by recalling that the non-BPS D-particles of type I are dual to perturbative non-BPS states of the $`SO(32)`$ heterotic string which also have gravitational and gauge interactions among themselves. These can be computed using standard perturbative methods and if one takes into account the known duality relations and renormalization effects, one can explicitly check that they agree with the expressions (27) and (28). This agreement provides further dynamical evidence of the heterotic/type I duality beyond the BPS level. |
warning/0003/cs0003059.html | ar5iv | text | # SATEN: An Object-Oriented Web-Based Revision and Extraction Engine
## General Information
SATEN<sup>1</sup><sup>1</sup>1SATEN a Sagacious Agent for Theory Extraction and revisioN is available at http://cafe.newcastle.edu.au/saten is an object-oriented web-based extraction and belief revision engine. It runs on any computer via a Java 1.1 enabled browser such as Netscape 4. SATEN performs belief revision based on the AGM approach (?). The extraction and belief revision reasoning engines operate on a user specified ranking of information. One of the features of SATEN is that it can be used to integrate mutually inconsistent commensuate rankings into a consistent ranking.
## General Description of the System
SATEN has two reasoning engines: a revision engine and an extraction engine. Both engines operate on a user-specified ranking of information.
The process of extraction involves the recovery of a consistent ranking from an inconsistent ranking. Several strategies are available for this process. Extraction can be seen as a generalisation of Brewkaโs work on subtheories (?). It is also related to Pooleโs work (??).
Revision models the incorporation of new information (a belief) into a knowledge base. The main difficulty in modeling this process is that the incoming information may be inconsistent with the agentโs knowledge base. Consequently, the agent must determine the information to be relinquished before the new information can be accepted, if the agent is to maintain a consistent knowledge base. There are several constructive definitions of belief revision based on various preference relations over the agents knowledge. These preference relations attempt to capture the relative importance of information with respect to change. When an agent is forced to give up beliefs in order to accept some new information then, if presented with a choice, a response would be to choose to retain more important information, like system laws and integrity constraints, rather than contingent facts and default facts.
SATEN is capable of:
* Theory Extraction
* Iterated Belief Revision
* Nonmonotonic Reasoning
* Possibilistic Reasoning
* Hypothetical Reasoning
Using combinations of these capabilities SATEN can perform information integration/fusion on mutually inconsistent commensurate knowledge bases and can calculate Spohnian reasons (?).
SATEN currently implements the reasoning strategies below based on a user specified ranking of beliefs (logical formulae). A ranking can be specified as mapping from beliefs to \[1,2,3,โฆ\] in which case the most important information is given the lowest rank, i.e. rank is inversely proportional to importance. Alternatively, beliefs can be ranked using degrees of belief between 0 and 1 - the higher the degree the more important the information, in particular tautologies are assigned 1, whilst inconsistencies and nonbeliefs are given 0. SATEN can derive an ordinal ranking from a set of specified degrees automatically, by toggling one of the menu settings.
* Standard Adjustment (???) is based on the standard epistemic entrenchment construction (?) for Belief Revision.
* Maxi-Adjustment is based on maximal inertia (?). Maxi-adjustment proceeds from the top of the ranking and moves down it rank by rank. At each rank it deletes all the beliefs that are inconsistent with other beliefs at that rank and above.
* Hybrid Adjustment is a combination of standard adjustment and maxi-adjustment (?). Adjustment is computationally โeasierโ than maxi-adjustment, however the main shortcoming of adjustment is that it maintains beliefs only if there is an *explicit reason to keep them*, whilst maxi-adjustment removes beliefs only if there is an *explicit reason to remove them*. Hybrid adjustment performs an adjustment which computes the core beliefs to be retained, and then it performs a maxi-adjustment which is used to regather as many beliefs as possible that were discarded during the adjustment phase.
* Global Adjustment is similar to a maxi-adjustment except that it takes all beliefs in the ranking into consideration when computing the minimally inconsistent subsets, instead of proceeding rank by rank. It removes the least ranked beliefs โcausingโ the inconsistency.
* Linear Adjustment is similar to maxi-adjustment except that it removes all the beliefs at ranks which contain incosisitencies with beliefs above them (?). IF there is only one belief at each rank, then linear and maxi-adjustment are identical.
* Quick Adjustment is identical to maxi-adjustment except that rather than removing all beliefs it randomly chooses a minimal number of beliefs from each set of inconsistent beliefs in this way the inconsistency is removed.
A number of examples are available via a dropdown menu which highlight the different behaviour of the strategies.
## Methodology
An inherent difficulty that arises when trying to model world states using the predicate calculus is that of consistency. Under the rules of formal logic, anything is derivable from a contradiction, hence it is imperative that consistency of the model be maintained at all times so that the agent acting on that model is able to make sound decisions. However, it is not uncommon in the realm of belief-based reasoning for newly acquired information to be inconsistent with the current set of beliefs. For instance, suppose that an agent believes that Tweety is a bird, and that all birds can fly. This agent subsequently discovers that, Tweety is a penguin, and knowing that penguins canโt fly, will now have to deal with the fact that this new information is inconsistent with what it already believes.
It is clear that when this sort of situation arises, the agent will need to select a set of beliefs to be removed from its knowledge base in order to maintain consistency in its knowledge base when the new belief is inserted, thus allowing it to continue to reason on the basis of its revised knowledge base. However, the question arises: which beliefs should be removed? Even in the above example, the removal of any subset of the beliefs listed will make the new information that Tweety is a penguin consistent, so the question of which belief or set of beliefs to remove has no clear answer.
Alchourrรณn, Gรคrdenfors and Makinson (???) investigated a set of rationality postulates regarding theory change operators. These postulates appear to capture much of what is required of an ideal rational system of theory change. In particular, they developed three primary classes of theory change operators: expansion, contraction and revision. Expansions are deterministic, and can be expressed set-theoretically, but it is well known that the logical properties of a body of information are not strong enough to uniquely determine a revision or contraction operator. Consequently, some sort of preference order needs to be imposed on the beliefs in the agentโs knowledge base. This approach to Belief Revision has come to be known as the AGM paradigm.
Two common models used to impose the preference order required by the AGM paradigm on systems of beliefs are epistemic entrenchment orderings (?) and systems of spheres (?). An epistemic entrenchment ordering ranks the sentences in the knowledge base, whereas a system of spheres is a total preorder of the set of possible world states. SATEN uses the epistemic entrenchment representation and uses the services of a full first order theorem prover implemented in Java (?). Revision of a system of spheres can be implemented via a standard Database Management System (?).
The implementation of revision in SATEN is steeped in the AGM paradigm (?). Gรคrdenfors and Makinsonโs (?) showed that an epistemic entrenchment can uniquely determine how an agent will react to the pressures of impinging information. Their construction is an elegant theoretical result, and to use it as the foundation for a computer-based implementation of belief revision the following problems need to be addressed.
* Epistemic entrenchments order all the beliefs in the underlying language. The underlying language may be infinite in size, therefore a finite representation for an epistemic entrenchment is required.
* A new knowledge base, rather than a new epistemic entrenchment ordering is constructed. However from a practical point of view developing a new entrenchment ranking is imperative because an implementation must have the capacity to perform iterated revision.
An implementation of belief revision must propagate an epistemic entrenchment ordering (or more precisely its finite representation). SATEN overcomes the first problem using a finite partial entrenchment ranking (??) as the underlying representation scheme, and it overcomes the second problem using transmutations (???).
## Applying the System
The user specifies a prioritised knowledge base. We have found that for most applications the user-specified ranking can be otained in a natural way during the analysis and design phase (?).
The following guidelines for designing a ranking were developed in (?):
* Important information should be made explicit in the ranking
* Information should be represented in its simplest logical form
* Maximize the number of ranks, and minimise the number of beliefs at each rank for better performance
* Rankings should be as irredundant as possible
* Subsumption at the same rank should be avoided
* User specifies reasons
* Check the contraposition of reasons does not lead to undesired behaviour
SATEN can revise first order theory bases. It uses the services a Java first order theorem prover VADER (?) to determine consequence relations, and hence takes worst-case exponential time in the number of formulae in the knowledge base to compute even propositional revision. However, given Horn input clauses, the system will use a Linear Descent algorithm for its deductions, resulting in linear time efficiency. A number of strategies have been implemented in this system : standard adjustment, maxi-adjustment, hybrid adjustment, global adjustment, linear adjustment, and quick adjustment.
Theory extraction is related to belief revision and uses an extraction operator is a map between epistemic entrenchment orderings which, given an inconsistent ranking, extracts from it some (in some sense) maximal consistent ranking/theory base. Whilst every extraction operator gives rise rather naturally to a corresponding revision operator, it remains unknown whether the converse is true. When SATEN performs a revision, the internal behaviour is to add the new belief to the base at an appropriate rank, and then to perform a Theory Extraction, before normalising the beliefs to achieve the revised theory base. The process of normalisation converts any ranking into a partial entrenchment ranking.
SATEN is capable of performing both Belief Revision and Theory Extraction on theory bases, and of determining the degree of an arbitrary wff in itโs theory base, and can normalize a ranked theory base. It is also capable of automatically ranking the beliefs in the theory base โ the user has the option to view the total order as a ranking (with the (0,1)-values of the entrenchments suppressed) although the rankings are always represented internally with (0,1)-values. Finally, the user may specify whether new beliefs added to the base without a rank should, by default, be placed at the top of the ranking, or at the bottom. This feature can be used when the rank of new information is unknown. The most conservative approach would place the new information without rank at the both of the ranking.
### Input Format
SATEN accepts any Predicate Calculus wffโs as input. Variables in the predicate calculus must begin with a capital letter, and contain only alphanumeric symbols and underscores (\_). Constants have the same format as variables, but must begin with a lower case letter. Predicate and function names must comply with the same format as variables and constants respectively, but must be followed by a parenthesized list of arguments separated by commas.
Input should be in fully parenthesised logic syntax and should not contain spaces:
* Negation is a minus sign (-).
* Implication is an arrow composed of a minus sign and a greater than sign ($`>`$). Note no space between the two symbols.
* Conjunction is an ampersand (&).
* Disjunction is a pipe ($`|`$).
* The universal quantifier is an asterisk (\*).
* The existential quantifier is an exclamation mark (!).
* Constants must begin with lower-case letters.
* Variables with upper-case letters.
* Functions start with lower-case.
* Propositional statements start with lower-case.
* First Order Predicates begin with upper-case.
Expression grouping is with parentheses, as is function (and predicate) argument list demarkation. User input should not include any double underscores (\__) as these are used by the program to tag system-generated variables and skolem functions.
Example input format: \*X(Psychopathic(X)$`|`$ Emotional(X))
-null
hopes&dreams
\*X(!Y(Mother(X,Y)))
!Z(EatsChocolate(Z)$`>`$Happy(Z))
\*Y(-Income(Y)$`>`$-Loan(Y))
We can encode โ โFor all $`X`$ there exists a $`Y`$ such that $`Y`$ is the mother of $`X`$, and for all $`Y`$, if $`Y`$ is the daughter of Benโs maternal grandmother, then $`Y`$ is Benโs mother, or $`Y`$ is Benโs aunt.โ using the following sentence:
$$(X(!Y(M(X,Y)))\&(Y(D(MGM(ben),Y)>$$
$$(M(ben,Y)|A(ben,Y)))$$
where $`X`$ and $`Y`$ are variables, $`M(X,Y)`$ is a predicate of two variables with name $`M`$, $`MGM(X)`$ is a function of one variable with name $`MGM`$, and $`ben`$ is a constant.
### Strategies
The standard adjustment strategy finds the largest cut of the inputs to SATEN that is consistent with the incomming information, and keeps this as the result of the revision. This is quite fast, but removes alot of beliefs from the theory base in comparison to the other strategies because as shown in (?) the independence of beliefs must be explicitly specified in the ranking.
The maxi-adjustment strategy looks at each individual rank in the belief set, and removes the union of all subsets $`S`$ of the beliefs on that rank having the properties
1. $`S`$ is inconsistent with the union of the incomming information and all the information ranked higher in the base than $`S`$; and
2. no proper subset of $`S`$ satisfies $`1`$.
This strategy keeps many of the beliefs in the theory base, but is also quite slow as it has to examine all subsets of each rank. Performance is improved if the agent is more discerning. our algorithm is anytime (?) so the longer the agent has to revise the better the result.
The hybrid adjustment strategy represents a combination of the standard adjustment and maxi-adjustment strategies. When revising with $`a`$, hybrid adjustment first finds every belief $`b`$ in the base such that $`a|b`$ is in the base, and is at the same rank as $`a`$ (this is the adjustment step โ it is removing every belief that โexplicitlyโ fails to be a consequence of the cut above $`a`$), and then performs a maxi-adjustment on what is left. So in other words maxi-adjustment is used to recoup as much of the ranking as the time set aside for the revision permits.
This strategy is comparable to maxi-adjustment with regard to the number of beliefs retained, and though somewhat faster on average, has the same worst-case time expenditure as the maxi-adjustment procedure.
The global adjustment strategy behaves exactly as maxi-adjustment does, but acting under the assumption that all beliefs have the same rank. This strategy is quite slow, and is not as parsimonious in what it removes as maxi-adjustment, but is highly intuitive with respect to the set of beliefs removed (without reference to their ranks). It comes as close as possible to eliminating the need for a ranking.
The linear adjustment strategy (?)looks at the successive cuts of the theory-base and, at every rank at which the cut is inconsistent, removes the entire rank from the base. This is quite time-efficient, and keeps more beliefs in the base than standard adjustment, but, unlike hybrid adjustment and maxi-adjustment, is fairly ad-hoc about the beliefs to be kept in the theory base.
The quick adjustment strategy is an attempt to capture the judiciousness of maxi-adjustment without the associated time expenditure. Once a given rank of beliefs is discovered to cause a contradiction, quick adjustment begins adding beliefs from the left end of that rank until a contradiction is discovered. It then removes these beliefs, one at a time, keeping track of which ones eliminate the contradiction when they are removed. It is these โculpritโ beliefs that are then removed from the theory base. This process is then iterated. This strategy has the advantage of improving performance by reducing the amount of processing at each rank, and also emulates the output of maxi-adjustment much of the time. However, its behaviour depends upon the order of the inputs on a given level of the theory base, and hence is somewhat ad hoc. The system could be easily extended to generate more than one ranking and used to perform credulous reasoning. Instead of randomly choosing an alternative, all alternative revised rankings could be created.
### Other Options
SATEN implements three additional (optional) strategies that interact with the various Belief Revision strategies outlined above. These strategies are subsumption removal, recovery, and nuclear revision. Of these three, the second two are appropriate with any belief revision strategy, but subsumption removal cannot be used in conjunction with the standard adjustment strategy. This is the only type of nuclear revision currently implemented. Other approaches like attributing half-lifes to individual beliefs will be available in the next version of SATEN.
Using subsumption removal ensures that any strategy which must choose from a set of beliefs to remove will preferrentially remove beliefs that are subsumed by the incomming belief. For example, maxi-adjustment will look amongst the set of beliefs it was going to remove for the subset that are subsumed by the incomming belief. If there are none, it will proceed as before, but if there are some, it will throw them away first, and then check again for consistency. If the result is still not consistent, the maxi-adjustment algorithm is run again on the remaining beliefs.
The recovery option ensures that the recovery postulate is satisfied by the revisions. This is achieved by, rather than removing a belief $`b`$, replacing it with the disjunction $`b|a`$ where $`a`$ is the new belief. The only time we do not do this is if $`b|a`$ is a tautology.
The concept of nuclear revision models the fact that beliefs that are not constantly reinforced tend to become less substantial to an intelligent agent. When nuclear revision is enabled, the user will be asked for a half-life. This is the constant factor by which the entrenchment of every belief in the base drops after each operation on the theory base. Thus, if the half-life were set at $`05`$, then the entrenchment of each belief in the theory base would be halved after each operation.
### Examples
SATEN has a substantial example menu intended to indicate the correct input form for SATEN, and demonstrate SATENโs capabilities, but also to illustrate the differences between the various belief revision strategies that SATEN implements. These examples are broken into three categories: propositional, predicate, and contrast of strategies.
The first of these indicates the correct input and behaviour of SATEN when presented with propositional inputs only. Under these conditions, SATEN tends to be somewhat more efficient than in the predicate case for the reasons outlined in the VADER hyperbook in the section on performance issues.
The second set of examples indicate SATENโs behaviour with the more general input form of the Predicate Calculus, specifically, its behaviour when quantifiers and variables are involved. The differences are essentially at the theorem prover level, and are opaque to SATEN itself.
The third set of examples are designed to indicate differences in behaviour between the different revision strategies. This is done for two reasons: the first is to highlight the salient points that make each strategy desirable or undesirable in comparison to the others; the second, which is, to some degree subsumed by the first, is to demonstrate that there actually is a difference between each pair of strategies, and that these differences do not require overly sophisticated examples to demonstrate โ in particular, propositional calculus is sufficient, and it is possible to get pairwise distinct behaviour from the strategies on an example with just nine beliefs and four distinct ranks. |
warning/0003/cond-mat0003168.html | ar5iv | text | # Low field negative magnetoresistance in double layer structures
## I Introduction
Transitions between 2D layers is one of fundamental features of double layer structures. It changes the quantum corrections to the conductivity, especially in a magnetic field parallel to the layers.
It is well known that the interference of electron waves scattered along closed trajectories in opposite directions (time-reversed paths) produces a negative quantum correction to the conductivity. An external magnetic field ($`๐`$) gives the phase difference between pairs of time-reversed paths $`\phi =2\pi (\mathrm{๐๐})/\mathrm{\Phi }_0`$, where $`\mathrm{\Phi }_0`$ is the quantum of magnetic flux, $`๐`$ is the area enclosed, and thus destroys the interference and results in negative magnetoresistance.
In case of a single 2D layer the influence of a magnetic field is strongly anisotropic because all the paths lie in one plane. The magnetoresistance is maximal for $`๐๐ง`$, where $`๐ง`$ is the normal to 2D layer. When a magnetic field lies in the 2D layer plane, $`๐๐ง`$, the scalar product $`(\mathrm{๐๐})`$ is zero, i.e. the magnetic field does not destroy the interference, and the negative magnetoresistance is absent in this magnetic field orientation.
In coupled double layer structures, the tunneling between layers gives rise to the closed paths where an electron moves initially over one layer then over another one and returns to the first layer. For this paths the product ($`\mathrm{๐๐}`$) is non-zero for any magnetic field orientation and hence the negative magnetoresistance has to appear for $`๐๐ง`$ as well.
The magnetic field dependence of the negative magnetoresistance is determined by the statistics of closed paths, namely, by the area distribution function, $`W(๐)`$, and area dependence of the average length of closed paths, $`\overline{L}(๐)`$. Just these statistic dependencies have been studied in single 2D layer structures by analysis of negative magnetoresistance measured at $`๐๐ง`$.
The role of inter-layers transitions in weak localization and negative magnetoresistance for $`๐๐ง`$ for multilayer structures (superlattices) was discussed in Ref. . The closely related problem concerning the role of inter-subbands transitions in quasi-two dimensional structures with several subbands occupied was theoretically studied in Ref. .
In this paper we present the results of investigations of the negative magnetoresistance in double layer GaAs/InGaAs structure for different magnetic field orientations. We obtain the area distribution functions and area dependencies of the average lengths of the closed paths using the approach developed in Refs. . These functions are compared with those obtained from the computer simulation of carrier motion when inter-layers transitions are accounted for. Close agreement shows that just the inter-layers transitions determine the negative magnetoresistance in coupled double layer structures.
## II experimental results
The double well heterostructure GaAs/In<sub>x</sub>Ga<sub>1-x</sub>As was grown by Metal-Organic Vapor Phase Epitaxy (MOVPE) on semi-insulator GaAs substrate. The heterostructure consists of a 0.5 $`\mu `$m-thick undoped GaAs epilayer, a Si $`\delta `$-layer, a 75 ร
spacer of undoped GaAs, a 100 ร
In<sub>0.08</sub>Ga<sub>0.92</sub>As well, a 100 ร
barrier of undoped GaAs, a 100 ร
In<sub>0.08</sub>Ga<sub>0.92</sub>As well, a 75 ร
spacer of undoped GaAs, a Si $`\delta `$-layer and 1000 ร
cap layer of undoped GaAs. The samples were mesa etched (cut??) into standard Hall bridges. The measurements were performed in the temperature range $`1.54.2`$ K at low magnetic field up to $`0.4`$ T with discrete $`10^4`$ T for two orientations: the magnetic field was perpendicular ($`๐๐ณ`$) and parallel ($`๐๐ฑ`$) to the structure plane (see insert in Fig. 1). Additional high field measurements were also made to characterize the structure. It has been found that in the structure investigated the conductivity is determined by the electrons in the wells. Their densities have been determined from the Fourier analysis of the Shubnikov-de Haas oscillations and consist of $`4.5\times 10^{11}`$ cm<sup>-2</sup> and $`5.5\times 10^{11}`$ cm<sup>-2</sup> in different wells. The Hall mobility was about $`\mu 4200`$ cm$`{}_{}{}^{2}/`$ (V $`\times `$ sec).
The magnetic field dependencies of in-plane magnetoconductance
$$\mathrm{\Delta }\sigma (B)=\sigma (B)\sigma (0)=1/\rho (B)1/\rho (0)$$
(1)
at magnetic field perpendicular ($`\mathrm{\Delta }\sigma (B_z)`$) and parallel ($`\mathrm{\Delta }\sigma (B_x)`$) to the structure plane are presented in Fig. 1. One can see that the negative magnetoresistance is observed for both magnetic field orientations and, in contrast to the case of single layer structures, the effects are comparable in magnitude. Analysis of behaviour of the conductivity in a wide range of temperatures ($`1.5<T<20`$ K) and magnetic fields ($`B<6`$ T) shows that at $`B<0.40.5`$ T the main contribution to the negative magnetoresistance comes from the interference correction. In this case the magnetic field dependence of the negative magnetoresistance is determined by the statistics of the closed paths.
Let us apply the method proposed in Refs. to analysis of negative magnetoresistance in the double layer structure. Using formalism presented in Section II of Ref. one can write the expression for conductivity of double layer structure with identical layers for two magnetic field orientations as follows
$`\sigma (B_i)`$ $`=`$ $`\sigma _0+\delta \sigma (B_i)`$ (2)
$`=`$ $`\sigma _04\pi l^2G_0{\displaystyle _{\mathrm{}}^{\mathrm{}}}dS_i\{W(S_i)`$ (4)
$`\mathrm{exp}({\displaystyle \frac{\overline{L}(S_i)}{l_\phi }})\mathrm{cos}\left({\displaystyle \frac{2\pi B_iS_i}{\mathrm{\Phi }_0}}\right)\},i=x,z.`$
Here, $`G_0=e^2/(2\pi ^2\mathrm{})`$, $`\sigma _0`$ is the classical Drude conductivity, $`l=v_F\tau `$, $`l_\phi =v_F\tau _\phi `$, $`v_F`$ is the Fermi velocity, $`\tau `$ and $`\tau _\phi `$ are the momentum relaxation and phase breaking time, respectively. The value of $`\overline{L}`$ is the function not only of $`S`$ but $`l_\phi `$ as well. It was defined in Ref. by Eq.(6). It should be mentioned that Eq. (2) is valid at low enough probability of inter-layers transitions.
Thus, for the magnetic field perpendicular to the structure plane ($`๐=(0,0,B_z)`$) the magnetoresistance is determined by $`z`$-component of $`๐`$ only and for parallel magnetic field ($`๐=(B_x,0,0)`$) it is determined by $`x`$-component of $`๐`$.
One can see from Eq. (2) that the Fourier transform of $`\delta \sigma (B)/G_0`$
$`\mathrm{\Phi }(S_i,l_\phi )`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Phi }_0}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐B_i{\displaystyle \frac{\delta \sigma (B_i)}{G_0}}\mathrm{cos}\left({\displaystyle \frac{2\pi B_iS_i}{\mathrm{\Phi }_0}}\right)=`$ (5)
$`=`$ $`4\pi l^2W(S_i)\mathrm{exp}\left({\displaystyle \frac{\overline{L}(S_i)}{l_\phi }}\right),`$ (6)
carriers an information on $`W(S_i)`$ and $`\overline{L}(S_i)`$. Because the value of $`l_\phi `$ tends to infinity when temperature tends to zero, the extrapolation of $`\mathrm{\Phi }(S_i,l_\phi )`$-vs-$`T`$ curve to $`T=0`$ gives the value of $`4\pi l^2W(S_i)`$. The ratio $`\overline{L}(S_i)/l_\phi `$ for given $`S_i`$ can be then obtained as $`\mathrm{ln}(4\pi l^2W(S_i))\mathrm{ln}(\mathrm{\Phi }(S_i,l_\phi ))`$.
The value $`\mathrm{\Delta }\sigma (B)=\sigma (B)\sigma (0)`$, not $`\delta \sigma (B)`$, is experimentally measured. It is clear from Eqs. (1) and (2) that $`\delta \sigma (B)=\sigma (0)\sigma _0\mathrm{\Delta }\sigma (B)`$. To obtain $`\delta \sigma (B)`$, we assume that the Drude conductivity $`\sigma _0`$ is equal to the conductivity at T=20 K, when the quantum corrections are small. Notice that the final results are not sensitive to the value of $`\sigma _0`$ practically. Obtaining of the distribution function $`W(S_x)`$ from the experimental $`\delta \sigma (B_z)`$ dependencies for the structure investigated is illustrated by Fig. 2. In left panel the Fourier transforms of $`\delta \sigma (B_x)`$ measured at different temperatures are presented. Right panel shows how the $`\mathrm{\Phi }`$-vs-$`T`$ data have been extrapolated to $`T=0`$. The area distribution function $`W(S_z)`$ has been obtained from the analysis of $`\delta \sigma (B_z)`$ curves in a similar way.
The results of data processing described above are presented in Fig. 3a. As is seen the $`4\pi l^2W(S_z)`$ dependence is close to $`(2S_z)^1`$ for $`S(0.35)\times 10^{10}`$ cm<sup>2</sup>. The analogous behaviour of the area distribution function was obtained for single 2D layer in Ref. . The behaviour of $`W(S_x)`$ significantly differs from that of $`W(S_z)`$. In particular, the $`W(S_x)`$ curve shows a much steeper decline for $`S>0.8\times 10^{10}`$ cm<sup>2</sup> . Other feature of the statistics of the closed paths in double layer structure is the fact that for given $`S`$ the values of $`\overline{L}(S_x)`$ are significantly larger than $`\overline{L}(S_z)`$ (Fig. 3b).
Qualitatively these peculiarities of the statistics of closed paths in double layer structures can be understood if one considers how trajectories with large enough length, $`Ll/t`$, look. They are isotropically smeared over $`xy`$-plane for the distance $`\sqrt{Ll}`$, their extended area in this plane is $`s_zLl`$. In $`yz`$-plane they have size $`\sqrt{Ll}`$ in $`y`$-direction and $`Z_0`$ (where $`Z_0`$ is inter-layer distance) in $`z`$-direction. So, the extended area in $`yz`$-plane is $`s_xZ_0\sqrt{Ll}`$. Thus, closed trajectories have significantly larger $`s_z`$ than $`s_x`$, and the $`s_z/s_x`$ ratio increases with increasing $`s`$. It is clear that the behaviour of enclosed areas $`S_z`$, $`S_x`$ is analogous. Therefore, for $`S_x=S_z`$ the inequality $`W(S_z)>W(S_x)`$ is valid. The average length of the trajectories $`\overline{L}(S_x)`$ therewith is greater than $`\overline{L}(S_z)`$.
As was shown in Ref. the distribution function of closed paths, the area dependence of average length of closed paths and weak localization magnetoresistance can be obtained by computer simulation of a carrier motion over 2D plane.
## III Computer simulation
The model double layer system is conceived as two identical plains with randomly distributed scattering centers with a given total cross-section. Every plane is represented as a lattice $`M\times M`$ with lattice parameter $`a`$. The scatterers are placed in a part of the lattice sites with the use of a random number generator. We assume that a particle moves with a constant velocity along straight lines which happen to be terminated by collisions with the scatterers. After every collision the particle has two possibilities: it passes from one plane to another with a probability $`t`$ and moves over the second plane or it remains in the plane with probability $`(1t)`$, changing the motion direction only. If the trajectory of the particle passes near the start point at the distance less than $`d/2`$ (where $`d`$ is a prescribed value, which is small enough), it is perceived as being closed. The projections of enclosed algebraic area is calculated according to
$`S_z={\displaystyle \underset{j=1}{\overset{N1}{}}}{\displaystyle \frac{y_{j+1}+y_j}{2}}(x_{j+1}`$ $``$ $`x_j)+`$ (7)
$`+{\displaystyle \frac{y_N+y_1}{2}}(x_N`$ $``$ $`x_1),`$ (8)
$`S_x={\displaystyle \underset{j=1}{\overset{N1}{}}}{\displaystyle \frac{y_{j+1}+y_j}{2}}(z_{j+1}`$ $``$ $`z_j)+`$ (9)
$`+{\displaystyle \frac{y_N+y_1}{2}}(z_N`$ $``$ $`z_1),`$ (10)
where $`N`$ is number of collisions for given trajectory, $`x_j`$, $`y_j`$, $`z_j`$ stand for coordinates of $`j`$-th collision, $`z_j`$ takes the value $`0`$ or $`Z_0`$. Otherwise the simulation details are analogous to them described in Ref. for the case of single 2D layer system.
All the results presented here have been obtained using the parameters: lattice dimension is $`6800\times 6800`$; the number of starts, $`I_s`$, is $`10^6`$; the total number of scatterers is about $`1.6\times 10^5`$; the scattering cross section is $`7`$; $`d=1`$; $`Z_0=18`$. The value of mean free path computed for such a system is about $`43\times a`$. If we suppose the value of $`a`$ equal to $`11`$ ร
, this model double layer system corresponds to the heterostructure investigated. Namely, the mean free path is equal to the value of $`l480`$ ร
, and the value of $`Z_0`$ is close to the distance between the centers of the quantum wells, $`200`$ ร
.
The area distribution functions obtained as the result of simulation with different inter-layers transition probabilities are presented in Fig. 4. Let us discuss at first the behaviour of $`W(S_z)`$ (Fig. 4a). For $`t=0`$, the $`4\pi l^2W(S_z)`$ curve corresponds to the area distribution function for single layer. For large $`S`$ this curve goes close to the $`S^1`$-dependence, which corresponds to the ideal 2D system in the diffusion regime. The deviation, which is evident for $`S_z<10^3a^2`$, is just due to the transition to the ballistic regime. It is obviously that for sufficiently large values of $`t`$, the probability of return to the start point has to be twice smaller than that for $`t=0`$. As is seen from Fig. 4a even the value $`t=0.1`$ is large enough in this sense: the corresponding curve is close to the $`(2S)^1`$-dependence practically in whole area range. This is because the long trajectories with large number of passes between layers give significant contribution to $`W(S_z)`$ starting from small areas, $`S_z>0.1l^2`$. For the intermediate value of $`t`$ ($`t=0.002,\mathrm{\hspace{0.33em}0.01}`$) the area distribution function is close to the $`S^1`$ function for small areas and tends to the $`(2S)^1`$ dependence for large ones.
The behaviour of $`W(S_x)`$ contrasts with that of $`W(S_z)`$ (Fig. 4b). At small $`S_x`$, $`W(S_x)`$ depends only weekly on $`S_x`$, whereas at large $`S_x`$ it decreases sharply when $`S_x`$ increases. Sensitivity of $`W(S_x)`$ to inter-layers transition probability depends on $`S_x`$ value. For small $`S_x`$ values, when the area distribution function is mainly determined by short closed paths with small number of inter-layers transitions, the value of $`W(S_x)`$ considerably increases with increasing $`t`$. For large $`S_x`$, i.e. for paths with large number of inter-layer transitions, $`W(S_x)`$ weakly depends on the transition probability.
Let us demonstrate how the magnetoresistance of our model 2D system changes with changing of the inter-layers transition probability. The theoretical $`\delta \sigma (B)`$ dependencies have been calculated by summing over the contributions of every closed path to the conductivity in accordance with following expression
$$\frac{\delta \sigma (B_i)}{G_0}=\frac{2\pi l}{I_sd}\underset{i}{}\mathrm{cos}\left(\frac{2\pi B_iS_i}{\mathrm{\Phi }_0}\right)\mathrm{exp}\left(\frac{l_i}{l_\phi }\right),$$
(11)
where $`l_i`$ is the length of $`i`$-th closed path. The results of calculation are presented in Fig. 5, where $`\mathrm{\Delta }\sigma (B_i)=\delta \sigma (B_i)\delta \sigma (0)`$ is plotted against $`B/B_t`$, $`B_t=\mathrm{}c/(2el^2)`$. As is seen the changes in magnetic field dependencies of negative magnetoresistance with change of inter-layers transition probability reflect the variation of area distribution functions. Indeed, $`\mathrm{\Delta }\sigma (B_z)`$ depends on $`t`$ slightly: maximal change is less than two times for decreasing $`t`$ up to zero, whereas $`\mathrm{\Delta }\sigma (B_x)`$ changes drastically. It decreases about hundred times, when the value of $`t`$ decreases from 0.1 to 0.002.
## IV Discussion
Let us compare the calculated area distributions with experimental data. One can see from Figs. 3a and 4, that the behaviour of calculated and experimental $`W(S_z)`$ and $`W(S_x)`$ dependencies is close qualitatively. As mentioned above, $`W(S_x)`$ depends on inter-layers transition probability significantly stronger than $`W(S_z)`$. Therefore we have estimated the transition probability comparing the calculated and experimental $`W(S_x)`$ curves. The most accordance has been obtained with $`t0.1`$ (see Fig. 3a). As seen from the figure, with this value of $`t`$ the calculated $`W(S_z)`$ dependence describes the experimental data well. Some quantitative inconsistency, especially for $`W(S_x)`$, is evident in Fig. 3a. We believe, this is result of crudity of model used. In particular, we supposed the identity of both 2D layers.
Let us turn now to magnetic field dependencies of negative magnetoresistance. To calculate $`\mathrm{\Delta }\sigma (B)`$, in addition to the inter-layers transition probability it is necessary to know the phase breaking length. Using the value of $`t=0.1`$ estimated above, we have found that the best agreement between theoretical and experimental $`\mathrm{\Delta }\sigma (B_i)`$ dependencies is obtained with $`l_\phi 3.4`$ and $`1.4`$ $`\mu `$m for T=1.5 and 4.2 K, respectively. The $`\mathrm{\Delta }\sigma (B_z)`$ and $`\mathrm{\Delta }\sigma (B_x)`$ dependencies calculated with these $`l_\phi `$ values practically coincide with those measured experimentally (see Fig. 1). It should be noted that these values of $`l_\phi `$ some differ from those obtained by fitting of the $`\mathrm{\Delta }\sigma (B_z)`$ curves to the Hikami expression: the fit gives $`l_\phi 4.8`$ and $`1.7`$ $`\mu `$m for T=1.5 and 4.2 K, respectively. The reason of this difference is that the Hikami formula was obtained for single 2D layer, and it is not suitable for analysis of negative magnetoresistance in coupled double layers structures.
Finally, knowing the values of $`t`$ and $`l_\phi `$ we are able to compare the calculated and experimental area dependencies of $`\overline{L}(S_x)`$ to $`\overline{L}(S_z)`$ ratio (Fig. 3b). It is seen that the experimental ratio is significantly larger than unity as well as calculated one. However, the experimental points lie somewhat below than calculated curve. We suppose that the main reason of such disagreement is dissimilarity of the layers in structure investigated.
After this paper has been prepared for publication, the paper by Raichev and Vasilopoulos on the theory of weak localization in double quantum wells is appeared in Condensed Matter e-Print archive. Let us apply this theory to our case. Using the formulae derived in Ref. , we have calculated $`\mathrm{\Delta }\sigma (B_i)`$ dependencies for our structure. These dependencies are represented in Fig. 1 by dashed curves. As is clearly seen, theory developed in Ref. describes our experimental results only in low magnetic fields. The reason is that the calculations in Ref. were carried out in the framework of diffusion approximation. It means that two conditions are met. The first condition is $`\tau \tau _\phi `$. For the structure investigated $`\tau /\tau _\phi =0.0140.035`$ for different temperatures, and this condition may be considered as fulfilled. According to the second condition, the magnetic field has to be low enough: $`BB_t`$ when $`๐๐ณ`$, or $`BB_tl/Z_0`$ when $`๐๐ฑ`$. In our case $`B_t0.14`$ T, $`l/Z_02.5`$ and hence the diffusion approximation is applicable, when $`B0.14`$ or $`0.35`$ T depending on the magnetic field orientation. It is in this range of magnetic field that the results of Ref. are close to our experimental data.
Our calculations are valid beyond the diffusion approximation and therefore they better describe the experimental results in whole magnetic field range, where weak localization correction to the conductivity is dominant.
## V conclusion
We have investigated the negative magnetoresistance in double layer heterostructures for different magnetic field orientations. The information about statistics of closed paths has been extracted from the analysis of temperature and magnetic field dependencies of conductivity. Significant difference in area distribution functions, $`W(S_x)`$, $`W(S_z)`$, and in average lengths of closed paths, $`\overline{L}(S_x)`$, $`\overline{L}(S_z)`$, has been found. In order to interpret the experimental results, we have investigated the statistics of closed paths and negative magnetoresistance using the computer simulation of the carrier motion with scattering over two 2D layers. Analysis of experimental and theoretical results unambiguously shows that in parallel magnetic field the negative magnetoresistance in double layer structures is determined by inter-layers transitions.
### Acknowledgments
This work was supported in part by the RFBR through Grant 00-02-16215, and the Program University of Russia through Grants 990409 and 990425. |
warning/0003/hep-th0003187.html | ar5iv | text | # Brane Dynamics in Background Fluxes and Non-Commutative Geometry
## 1 Introduction
Many important properties of D-branes can be understood in terms of open strings attached to the world-volume of the brane. At sufficiently low energies, the behavior of these open strings may be described through an effective field theory of the light modes from the open string spectrum. Typically, these field theories involve vector fields propagating on the brane along with a number of scalar fields that describe fluctuations of the brane in the transverse direction.
It was discovered more recently that the presence of a background B-field causes the world-volume of the brane to become non-commutative. The initial observation for branes on a 2-torus was clarified and generalized in a number of papers . One can see this non-commutative geometry of branes in their effective action. In fact, it naturally becomes a field theory on a non-commutative space so that the non-commutativity of the world-volume has direct influence on the dynamics of the brane.
Much of the existing work in this direction has been devoted to branes in flat space. A perturbative analysis along the lines of , on the other hand, shows that the quantization of world-volume geometries should be a much more general phenomenon which persists in the case of curved backgrounds. In fact, quantization of world-volumes appears to be inevitable in curved backgrounds because the string equation of motion enforce a non-vanishing NS 3-form field strength $`H`$ whenever the underlying space is not Ricci-flat. Since a non-zero field strength cannot possibly possess a vanishing 2-form potential $`B`$, branes in curved space are necessarily equipped with a non-zero B-field and hence they are quantized.
The first rigorous analysis of brane non-commutativity in a curved space, namely on a 3-sphere, was presented in , using the world-sheet approach to D-branes. In that work, we gave a prescription to extract non-commutative brane world-volumes from the data of a boundary CFT. This definition builds on the closed string ideas of (see also for examples) and is independent of notions from classical geometry, thus providing further evidence that brane non-commutativity is a general feature.
An exact conformal field theory description is available for certain maximally symmetric branes on an $`S^3`$. It was shown in that, classically, these branes are wrapped on conjugacy classes of $`\mathrm{SU}(2)S^3`$. The generic classes are 2-spheres in $`S^3`$, and due to the B-field they are quantized such that the algebra of โfunctions on the braneโ is a full matrix algebra or some (mildly non-associative) q-deformation thereof, depending on the size of the 3-sphere in which the branes are embedded. The size of the matrix algebra is determined by the radius of branes themselves. These world-volume algebras are also known as โfuzzy-spheresโ, see and references therein.
As these fuzzy spheres are now embedded into a full string theory with all its dynamical degrees of freedom, we may wonder about the field theory that describes the light open string modes living on the fuzzy sphere. Field theories with the expected field content have been described in the non-commutative geometry literature . In particular, Klimฤรญk has suggested a family of actions which are arbitrary linear combinations of non-commutative Yang-Mills and Chern-Simons terms. The string theory effective action is of this type but selects a particular member of this family for which the mass terms present in the non-commutative Yang-Mills and Chern-Simons actions cancel each other.
We will see that our field theories on the fuzzy sphere possess interesting classical solutions some of which have no analogue in effective field theories on flat branes. They describe stacks of spherical branes that evolve into superpositions of spherical branes with different radii. In particular, a stack of $`N`$ D0-branes (i.e. $`N`$ spheres of vanishing radius) can condense into a single D2-brane of radius $`rN`$. This dynamical generation of new world-volume dimensions occurs at the expense of a partial breaking the gauge symmetry.
The formation of extended objects from smaller constituents has of course been discussed extensively in the context of M-theory (see e.g. and references therein) and also for branes in constant RR-background fluxes . While the latter work takes the Born-Infeld action as a starting point, our analysis is based on an exact open string background given by a boundary CFT. In this way, we are also able to show that the effects in question are inherited from the braneโs non-commutativity.
This requires, however, to understand the brane dynamics beyond small fluctuation. The latter were studied recently in in order to establish the stability of branes on group manifolds. Many of the statements we will make below can easily be extended to other group manifolds, too. We restrict ourselves to the group SU(2) for simplicity and because the background $`\mathrm{SU}(2)S^3`$ is also an important ingredient of the CFT formulation of the Neveu-Schwarz 5-brane, see e.g. , such that our findings may be useful to study the geometry of D-branes in the presence of a stack of 5-branes. Similarly, our SU(2) WZW results are applicable in the study of branes on an AdS$`{}_{3}{}^{}\times S^3`$ string background, see e.g. .
The paper is organized as follows: In the next section we will briefly list a few facts about branes, B-fields and non-commutative geometry of branes in flat space, including a few important formulas that are rather convenient to prepare the ground for the following less standard discussion. Section 3 is mainly a review of the results in . It contains all the results we shall need for our computation of the effective action, which is carried out in Section 4. For simplicity, this will be illustrated in the bosonic theory first, but we will sketch how to extend the calculation to full supersymmetric string backgrounds like a NS 5-brane. It will turn out that the (bosonic part of the) low-energy effective action is the same as for bosonic strings on $`S^3`$. Section 5 contains a detailed analysis of classical solutions and their interpretation. Finally, we end with conclusions and some open problems.
## 2 Flat branes and non-commutative gauge theory
Before we show how to calculate low-energy effective actions for branes wrapping an $`S^2S^3`$, let us briefly review a few important steps of the analogous calculation in the standard case of branes in flat $`n`$-dimensional Euclidean space $`^n`$, or on a flat torus $`๐^n`$. Consider a D-brane which is localized along a $`p`$-dimensional hyper-plane $`V_p`$ in the target, with tangent space $`TV_p`$. The conformal field theory associated with such a Euclidean D-brane is defined on the upper half of the complex plane. It contains an $`n`$-component free bosonic field $`X=(X^\mu (z,\overline{z})),\mu =1,\mathrm{},n,`$ subject to Neumann boundary conditions in the directions along $`TV_p`$ and Dirichlet boundary conditions for components perpendicular to the world-volume of the brane. From the free bosons, one may obtain various new fields, in particular the vertex operators for the open string tachyons
$$V_k(x)=:\mathrm{exp}(ikX(x)):\text{ for all }kTV_p,$$
which can be inserted at any point $`x`$ on the real line. These operators correspond to the lightest states $`|k`$ within each sector of fixed momentum $`k`$. One can think of them as being assigned to the eigenfunctions $`e_k=\mathrm{exp}(iku)`$ of the momentum operator on the world-volume $`V_p`$ of the brane. To make this more explicit we will use the notation $`V[e_k](x)=V_k(x)`$.
When there is a magnetic field $`B=(B_{\mu \nu })`$ on the brane, the operator product expansion (OPE) of these U(1)-primaries reads
$$V[e_{k_1}](x_1)V[e_{k_2}](x_2)=\frac{1}{(x_1x_2)^{2\alpha ^{}k_1k_2}}e^{i\frac{\pi \alpha ^{}}{2}k_1^\mathrm{t}\mathrm{\Theta }k_2}V[e_{k_1+k_2}](x_2)+\mathrm{},$$
up to contributions involving descendant fields. In writing such OPEs of boundary operators we assume throughout the whole paper that their insertion points are ordered $`x_1>x_2`$. The product $`k_1k_2=G^{\mu \nu }k_{1,\mu }k_{2,\nu }`$ is to be taken with the inverse of the open string metric $`G`$ rather than of the original closed string metric $`g_{\mu \nu }`$. $`G^{\mu \nu }`$ can be read off from the standard expression $`h=\alpha ^{}G^{\mu \nu }k_\mu k_\nu =1/2G^{\mu \nu }(\sqrt{2\alpha ^{}}k_\nu )(\sqrt{2\alpha ^{}}k_\mu )`$ for the scaling dimension of the boundary primary fields $`V[e_k]`$. The relation of $`\mathrm{\Theta }`$ and $`G`$ to the closed string parameters $`B`$ and $`g=1`$ is given by
$$\mathrm{\Theta }=\frac{2}{BB^1}\text{ and }G=\frac{1}{1B^2}.$$
(2.1)
In a particular limit (โzero slope limitโ ) of the theory for which $`\alpha ^{}G^{\mu \nu }`$ becomes zero, the boundary fields have vanishing conformal dimension and the operator product expansions are no longer singular. Thus we obtain
$$V[e_{k_1}](x_1)V[e_{k_2}](x_2)=V[e_{k_1}e_{k_2}](x_2).$$
The product $`e_{k_1}e_{k_2}=\mathrm{exp}(i\frac{\pi \alpha ^{}}{2}k_1^\mathrm{t}\mathrm{\Theta }k_2)e_{k_1}e_{k_2}`$ is the Moyal-Weyl product of functions on $`V_p`$ (see for details). We can pass to a more universal form that does not make reference to the basis $`e_k`$ in the space of functions: If we introduce $`V[f]=\widehat{f}_kV[e_k]`$ with $`\widehat{f}_k`$ being the Fourier coefficients of $`f`$, we obtain
$$V[f](x_1)V[g](x_2)=V[fg](x_2).$$
This formula holds for arbitrary functions $`f,g`$ as long as we are in the zero slope limit. In order to compute the low-energy effective action, we need two further operator products involving the boundary current $`j^\mu (x)`$, namely
$`j^\mu (x_1)j^\nu (x_2)`$ $`=`$ $`{\displaystyle \frac{\alpha ^{}}{2}}{\displaystyle \frac{G^{\mu \nu }}{(x_1x_2)^2}}+\mathrm{},`$ (2.2)
$`j^\mu (x_1)V[f](x_2)`$ $`=`$ $`{\displaystyle \frac{\alpha ^{}G^{\mu \nu }}{x_1x_2}}V[i_\nu f](x_2)+\mathrm{}.`$ (2.3)
With the help of these equations, the $`n^{th}`$ order terms for massless fields in the LEEA are obtained from an $`n`$-point function of operators $`:j^\mu V[A_\mu ]:(x)`$ where $`A_\mu `$ is the (possibly non-abelian) gauge field, i.e. $`A_\mu (u)`$ are functions on the world-volume of the brane, which take values in $`\text{Mat}(N)`$ for a stack of $`N`$ identical branes. In the zero slope limit, the resulting action is given by a gauge theory on the non-commutative plane or torus ,
$$๐ฎ(A)_{\mathrm{flat}}=\frac{1}{4}d^puF_{\mu \nu }F^{\mu \nu }$$
where $`F_{\mu \nu }(A)=_\mu A_\nu _\nu A_\mu +i[A_\mu A_\nu ]`$ and where $``$ is the Moyal-Weyl product; the integration extends over the world-volume of the brane.
## 3 Branes on $`๐^\mathrm{๐}`$ and fuzzy spheres
Strings moving on a three-sphere are described by the SU(2) WZW model at level $`๐`$. Here, the radius $`R`$ of $`S^3`$ is related to the level by $`R\sqrt{\alpha ^{}๐}`$. Because $`S^3`$ is curved, the string equations of motion enforce the presence of a constant NS 3-form field strength $`H\mathrm{\Omega }/\sqrt{\alpha ^{}๐}`$ where $`\mathrm{\Omega }`$ is the volume form on the unit sphere.
The world-sheet swept out by an open string in $`S^3`$ is parametrized by a map $`g:\mathrm{H}\mathrm{SU}(2)`$ from the upper half-plane H into the group manifold SU(2)$`S^3`$. From this field $`g`$ one obtains Lie algebra valued chiral currents
$$J(z)=๐(g)g^1,\overline{J}(\overline{z})=๐g^1\overline{}g$$
as usual. There exists an exact conformal field theory description for maximally symmetric D-branes on SU(2), which are characterized by the gluing condition $`J(z)=\overline{J}(\overline{z})`$ along the boundary $`z=\overline{z}`$. The geometry of the associated branes is encoded in these gluing conditions . First of all, they force the endpoints of open strings to stay on conjugacy classes of SU(2). The generic such classes are 2-spheres, except for two cases where the classes degenerate into the single points $`e`$ resp. $`e`$, where $`e`$ is the unit element of SU(2). Moreover, the condition $`J=\overline{J}`$ selects a particular two form potential $`B`$ (โB-fieldโ) on the brane, which is given by
$$B=\frac{\text{Ad}(g)+1}{\text{Ad}(g)1}.$$
Here $`Ad(g)`$ denotes the adjoint action of the SU(2) on the tangent space of the brane. The existence of a non-vanishing B-field causes the brane world-volume to become non-commutative. From the previous analysis of branes in flat space we know that it is the antisymmetric tensor $`\mathrm{\Theta }`$ rather than the B-field itself (or its inverse) that describes the semi-classical limit of the braneโs non-commutative world-volume algebra. $`\mathrm{\Theta }`$ is computed from $`B`$ by means of the formula (2.1), leading to
$$\mathrm{\Theta }=\frac{2}{BB^1}=\frac{1}{2}\left(Ad(g)Ad(g)^1\right).$$
(3.4)
The semi-classical extension of the above analysis shows that, for fixed gluing conditions, only a finite number of SU(2) conjugacy classes satisfy a Dirac-type flux quantization condition . These โintegerโ conjugacy classes are the two points $`e`$ and $`e`$ along with $`๐1`$ of the spherical conjugacy classes, namely those passing through the points $`\mathrm{diag}(\mathrm{exp}(2\pi i\alpha /๐),\mathrm{exp}(2\pi i\alpha /๐))`$ for $`\alpha =\frac{1}{2},\mathrm{},\frac{๐1}{2}`$.
Our discussion of brane dynamics in the next section will be based upon a computation of the leading non-vanishing terms of the low-energy effective action around the point $`1/\alpha ^{}k=0`$. For this reason, let us briefly look at the classical geometry in the limit $`\alpha ^{}๐\mathrm{}`$ in which the three-sphere grows and approaches flat 3-space. If one parametrizes $`g`$ by $`X`$ taking values in the Lie algebra su(2), such that $`g1X`$, the expression for $`\mathrm{\Theta }`$ becomes $`\mathrm{\Theta }=\text{ad}(X)`$. This is the standard Kirillov bi-vector on the spheres in the algebra su(2)$`=^3`$. Consequently, the geometry of the limiting theory $`\alpha ^{}๐=\mathrm{}`$ is very close to the well-known situation of flat branes in a flat background with constant B-field, and we expect that the world-volume algebras of our branes in the WZW model will be quantizations of two-spheres equipped with their standard Poisson structure. According to Kirillovโs theory, all โintegerโ 2-spheres are quantizable and in one-to-one correspondence with irreducible representations of SU(2). Kirillovโs theory implies that โfunctionsโ on the quantized 2-spheres form matrix algebras $`\text{Mat}(2\alpha +1)`$ of size $`2\alpha +1`$ where $`\alpha `$ can be any of the numbers $`\alpha =0,1/2,1,\mathrm{}`$ In this way we have identified our branes with representations of SU(2) such that the braneโs label $`\alpha `$ \- which is proportional to its radius โ denotes the spin of the associated representation.
In addition to the geometric picture developed so far, we will now list a number of exact results from the boundary conformal field theory description of the associated quantum theory of open strings. Their derivation is explained in . The chiral fields of the SU(2) WZW model form an affine Kac-Moody algebra denoted by $`\widehat{\mathrm{SU}}(2)_๐`$. It is generated by fields $`J_a(z),a=1,2,3,`$ which possess the following operator product expansions
$$J_a(x_1)J_b(x_2)=\frac{๐}{2}\frac{\delta _{ab}}{(x_1x_2)^2}+\frac{if_{ab}^c}{(x_1x_2)}J_c(y)+\mathrm{}$$
(3.5)
where $`f`$ is the Levi-Civita tensor. The open string vertex operators of the D-brane theories we are dealing with are organized in representations of this symmetry algebra. For the brane with label $`\alpha =0,\frac{1}{2},\mathrm{},\frac{๐}{2}`$, there appear only finitely many such representations with integer angular momentum between $`J=0`$ and $`J=\text{min}(2\alpha ,๐2\alpha ^{})`$. Each representation $`J`$ contains ground states of lowest possible energy, which form $`(2J+1)`$-dimensional multiplets spanned by some basis vectors $`Y_m^J`$ with $`|m|<J`$. To these states we assign vertex operators $`V[Y_m^J](x)`$. They are similar to the tachyon vertex operators $`V[e_k]`$ for open strings in flat space. While the latter are associated with eigenfunctions of linear momentum $`k`$, our vertex operators $`V[Y_m^J]`$ carry definite angular momentum, and one may think of the elements $`Y_m^J`$ as spherical harmonics on the 2-sphere that forms the classical world-volume of the brane $`\alpha `$. Because of the quantization, there appears a cutoff on the angular momentum, i.e. $`J`$ is restricted by $`J\text{min}(2\alpha ,๐2\alpha ^{})`$. The operator product expansion of these open string operators was found to be of the form
$$V[Y_i^I](x_1)V[Y_j^J](x_2)=\underset{K,k}{}x_{12}^{h_Kh_Ih_J}\left[\begin{array}{ccc}I\hfill & J\hfill & K\hfill \\ i\hfill & j\hfill & k\hfill \end{array}\right]F_{\alpha K}\left[\begin{array}{cc}J\hfill & I\hfill \\ \alpha \hfill & \alpha \hfill \end{array}\right]V[Y_k^K](x_2)+\mathrm{}$$
(3.6)
where $`h_J`$ is the conformal dimension of $`V[Y^J]`$, $`[:::]`$ denote the Clebsch-Gordan coefficients of the group SU(2), and $`F`$ stands for the fusing matrix of the WZW-model. In the limit $`๐\mathrm{}`$, the fusing matrix elements approach the $`6J`$ symbols of the classical Lie algebra $`\mathrm{su}(2)`$. At the same time the conformal dimensions $`h_J=J(J+1)/(๐+2)`$ tend to zero so that the OPE (3.6) of boundary fields becomes regular as in a topological model.
For the limiting theory there exists a much more elegant way of writing the operator product expansion (3.6). To see this observe that SU(2) acts on the space of $`(2\alpha +1)\times (2\alpha +1)`$-matrices by conjugation with group elements evaluated in the $`(2\alpha +1)`$-dimensional representation. Under this action, the $`(2\alpha +1)^2`$-dimensional space $`\text{Mat}(2\alpha +1)`$ decomposes into a finite set of SU(2) modules of spin $`J=0,1,\mathrm{},2\alpha `$. This is precisely the list of representations which appears in the open string theory of our brane $`\alpha `$ when the level is sent to infinity. Hence, we can identify the ground states $`|Y_m^J`$ with matrices $`\mathrm{Y}_m^J\text{Mat}(2\alpha +1)`$. It is a remarkable feature of this identification that the multiplication of the matrices $`\mathrm{Y}_i^I`$ and $`\mathrm{Y}_j^J`$ turns out to encode the information about the operator products in our conformal field theory,
$$V[\mathrm{Y}_i^I](x_1)V[\mathrm{Y}_j^J](x_2)=V[\mathrm{Y}_i^I\mathrm{Y}_j^J](x_2).$$
The emergence of the matrix product $``$ in this relation was expected from our geometric analysis above. It shows that the world-volumes of branes in $`S^3`$ become fuzzy two-spheres in the quantum theory. We can now rewrite the operator product in a way which does no longer refer to a particular choice of a basis. In fact, for an arbitrary matrix $`๐ \text{Mat}(2\alpha +1)`$ with $`๐ =a_{Jm}\mathrm{Y}_m^J`$ we introduce $`V[๐ ]=a_{Jm}V[\mathrm{Y}_m^J]`$ and obtain
$$V[๐ _1](x_1)V[๐ _2](x_2)=V[๐ _1๐ _2](x_2)$$
(3.7)
for all $`๐ _1,๐ _2\text{Mat}(2\alpha +1)`$. This product allows us to compute arbitrary correlations functions of such vertex operators,
$$V[๐ _1](x_1)V[๐ _2](x_2)\mathrm{}V[๐ _n](x_n)=\mathrm{๐๐}(๐ _1๐ _2\mathrm{}๐ _n).$$
(3.8)
The trace appears because the vacuum expectation value is SU(2) invariant and the trace maps matrices to their SU(2) invariant component.
Let us close this section on the conformal field theory of branes on $`S^3`$ by displaying a further OPE, namely the one that involves a current $`J^a`$ along with the vertex operator $`V[\mathrm{Y}_j^J]`$,
$$J_a(x_1)V[\mathrm{Y}_j^J](x_2)=\frac{1}{x_1x_2}V[L_a\mathrm{Y}_j^J](x_2)+\mathrm{}\text{ with }L_a๐ :=[\mathrm{Y}_a^1\stackrel{}{,}๐ ]$$
(3.9)
for all matrices $`๐ \text{Mat}(2\alpha +1)`$. The operator $`L_a`$ expresses the action of the angular momentum on spherical harmonics.
## 4 Low-energy effective field theory
With the previous description of the world-sheet theory as a firm basis, we want to address the main aim of the paper, namely the construction of the effective action that describes D-branes wrapped on an $`S^2S^3`$ at low energies. The three-spheres, or rather the SU(2) WZW models, occur as internal spaces in (super) string backgrounds like AdS$`{}_{3}{}^{}\times S^3\times T^4`$ or in the world-sheet theory for the background with $`๐2`$ Neveu-Schwarz 5-branes, which provide a source for $`๐2`$ units of NS 3-form flux through a three-sphere surrounding their (5+1)-dimensional world-volume. The CFT description of such a NS 5-brane background involves a tensor product of 5+1 free bosons, of a linear dilaton and of an SU(2) WZW model . For our purposes, it will be sufficient to approximate the โexternalโ part $`M^7`$ of these targets by a flat space-time. At the moment, we furthermore restrict ourselves to bosonic strings, but it will turn out that the results for the low-energy effective action carry over the supersymmetric case.
Thus we consider stacks of $`N`$ D$`p`$-branes which span a certain flat world-volume in $`^7`$ and wrap a (fuzzy) two-sphere with label $`\alpha `$ in the $`S^3`$-component of space-time. The massless field of this brane configuration fall into two groups which we want to discuss separately.
First there are the vector bosons and scalars which are associated with the seven external directions transverse to the $`S^3`$. When we compute their contribution to the action, we have to use vertex operators $`:j^\mu V[A_\mu ]V[๐ ]:(x)`$. Here, $`\mu =0,\mathrm{},6`$ labels the coordinates transverse to the $`S^3`$, and $`A_\mu `$ is either a U($`N`$) gauge or a scalar field on the brane depending on whether the direction of $`\mu `$ is parallel or perpendicular to the brane. The third factor $`V[๐ ]`$ in the vertex operator corresponds to the ground states for the brane on $`S^3`$, which means that $`๐ `$ is a matrix of size $`2\alpha +1`$. The actual computation of the associated terms in the effective action proceeds exactly as for branes in flat space, except for the contribution from ground states in the WZW factor of the theory. But the evaluation of the latter is completely independent of the first seven directions, and our formula (3.8) shows that the WZW model contributes precisely in the same way as usual Chan-Paton factors do. Hence, the effective action for the fields $`A_\mu `$ appears as an ordinary non-abelian Yang-Mills-Higgs theory on a D$`(p2)`$-brane. The effect of the two dimensions which wrap the $`S^2S^3`$ is to replace U($`N`$) by the larger gauge group U($`N(2\alpha +1)`$). In particular, non-abelian gauge theories on the world-volume of a single brane appear for $`\alpha >0`$.
The discussion of the remaining three fields that come with the $`S^3`$ (scalars viewed from the external part of the brane world-volume) requires a lot more effort. To begin with, let us rescale the currents of the WZW model so as to bring their operator product expansions in a form similar to eqs. (2.2,2.3). As in the flat case, we extract the inverse of the open string metric from the conformal dimensions, here given by $`h_J=J(J+1)/(๐+2)`$. To leading order in $`1/๐`$, this yields the metric
$$G^{ab}=\gamma \delta ^{ab}\text{with }\gamma =2/๐.$$
(4.10)
For the Schwinger term in the OPE of WZW currents $`J_a`$ to resemble eq. (2.2), we introduce $`j^a=(\sqrt{2\alpha ^{}}/๐)J_a`$. The operator product expansions of these rescaled currents are
$`j^a(x_1)j^b(x_2)`$ $`=`$ $`{\displaystyle \frac{\alpha ^{}}{2}}{\displaystyle \frac{G^{ab}}{(x_1x_2)^2}}+\alpha ^{}{\displaystyle \frac{i๐ฟ_c^{ab}}{(x_1x_2)}}j^c(y)+\mathrm{}`$ (4.11)
$`j^a(x_1)V[๐ ](x_2)`$ $`=`$ $`{\displaystyle \frac{\alpha ^{}G^{ab}}{x_1x_2}}V[๐ซ_b๐ ](x_2)+\mathrm{}`$ (4.12)
where we also use a rescaled angular momentum operator $`๐ซ_a=L_a/\sqrt{2\alpha ^{}}`$. The structure constants $`๐ฟ`$ in the first line are obtained from the structure constants $`f`$ of the classical Lie algebra $`\mathrm{su}(2)`$ through $`๐ฟ_{ab}^c=f_{ab}^c/\sqrt{2\alpha ^{}}`$. Here and in the following, indices $`a,b,c\mathrm{}`$ on the objects $`๐ซ`$ and $`๐ฟ`$ are raised and lowered with the open string metric $`G`$. In particular, this implies that $`๐ฟ_c^{ab}=2f_c^{ab}/(\sqrt{2\alpha ^{}}๐)`$.
Let us focus on the WZW boundary fields $`:j^aV[๐ _a]:(x)`$ where $`๐ _a=a_a^{Jm}\mathrm{Y}_m^J`$ is an arbitrary set of matrices subject to the physical state condition $`๐ซ^c๐ _c=0`$; summation over the three directions $`a=1,2,3`$ tangent to the 3-sphere is understood. For simplicity, we shall omit the simple factors $`V[f]`$ that come with the seven transverse directions throughout most of our discussion. They are taken into account if we consider $`๐ _a`$ as matrix valued functions on the world-volume of the brane rather than constant matrices. $`๐ _a`$ may be regarded as three massless bosons that live on our fuzzy sphere (or the $`(p+1)`$-dimensional world-volume of the brane). In the limit $`๐\mathrm{}`$ they can also be viewed as massless fields depending on the $`p1`$ brane coordinates which are transverse to the $`S^3`$ because the conformal dimensions $`h_I`$ of the ground states $`|\mathrm{Y}_i^I`$ vanish as we send the level to infinity. When we want to describe a stack of $`N`$ of our $`\alpha `$-branes, the coefficients $`a_a^{Jm}`$ of the fields $`๐ _a`$ have to be $`N\times N`$ matrices and hence $`๐ _a\text{Mat}(N)\text{Mat}(2\alpha +1)`$. We stress, however, that the operators $`๐ซ_a`$ act exclusively on the second tensor factor and not on the Chan-Paton degrees of freedom.
We are now prepared to compute the effective action for the three fields $`๐ _a`$ to leading order in $`\alpha ^{}`$. As in the case of flat branes it turns out that only 3- and 4-point functions contribute to this order of the open string scattering amplitudes. For the 3-point function one finds the following expression
$`:j^aV[๐ _a]:(x_1):j^bV[๐ _b]:(x_2):j^cV[๐ _c]:(x_3)`$
$`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{(\alpha ^{})^2}{x_{23}x_{12}x_{31}}}[{\displaystyle \frac{x_{23}}{x_{12}}}G^{ab}๐ซ^{c,(1)}{\displaystyle \frac{x_{31}}{x_{12}}}G^{ab}๐ซ^{c,(2)}+{\displaystyle \frac{x_{31}}{x_{23}}}G^{bc}๐ซ^{a,(2)}{\displaystyle \frac{x_{12}}{x_{23}}}G^{bc}๐ซ^{a,(3)}`$
$`+{\displaystyle \frac{x_{12}}{x_{31}}}G^{ac}๐ซ^{b,(3)}{\displaystyle \frac{x_{23}}{x_{31}}}G^{ac}๐ซ^{b,(1)}i๐ฟ^{abc}]V[๐ _a]V[๐ _b]V[๐ _c]+O((\alpha ^{})^3).`$
Here, the superscript $`^{(j)}`$ means that the operator acts on the argument of the $`j^{th}`$ vertex operator in the correlation function. The terms involving $`๐ซ`$ arise from contracting two currents into a Schwinger term and acting on one of the primary fields with the third current. The structure constants $`๐ฟ^{abc}`$ appear upon contraction of two currents into the second term of eq. (4.11) and subsequent contraction of the resulting current with the third current in the correlator. In the bosonic situation, there are higher order terms which we did not spell out because they will not be needed below.
From formula (4), we recover the corresponding expression for the flat space theory by the replacements $`๐ซ^ai^a`$ and, more importantly, $`๐ฟ^{abc}0`$. The 3-point functions feel the non-abelian nature of the theory through the terms that involve the Lie algebra structure constants. In the 4-point functions, such terms are suppressed by at least one order of $`\alpha ^{}`$, and the leading $`O((\alpha ^{})^2)`$ contributions come from pairwise contractions of two currents into a Schwinger term. They are identical to the corresponding terms in the case of flat D-branes except for the replacement $`๐ซ^ai^a`$.
Before we can combine this information into some effective action, we need to consider the mass terms of the theory. Recall that the mass of open string modes is determined from the conformal dimensions in the internal sector of the theory through $`M^2=(h1)/\alpha ^{}`$. The modes we consider have conformal dimension $`h=h_J+1`$ so that, to leading order in $`1/๐`$,
$$M^2=\frac{1}{\alpha ^{}}h_J=\frac{J(J+1)}{\alpha ^{}๐}.$$
(4.14)
This relation shows that it is the combined limit $`\alpha ^{}0`$ and $`\alpha ^{}๐\mathrm{}`$ that is relevant for our low-energy considerations. The values in (4.14) constitute the spectrum of the operator $`\frac{2}{๐}_a๐ซ_a๐ซ_a=๐ซ_a๐ซ^a`$ so that the formula for $`M^2`$ results in the following mass term for the Lagrangian of our theory
$$๐ฎ_0=\frac{1}{2}\mathrm{๐๐}\left(๐ _a๐ซ_b๐ซ^b๐ ^a\right)=\frac{1}{2}\mathrm{๐๐}\left(๐ซ_b๐ _a๐ซ^b๐ ^a\right).$$
This is the free theory to which we add the interaction terms encoded in the 3- and 4-point functions. They are obtained as usual, with the help of the physical state condition $`๐ซ^a๐ _a=0`$, by multiplication with ghost propagators and after summation over all permutations of the insertion points $`x_i`$. The result is
$`๐ฎ^{(3)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{๐๐}\left(๐ซ_a๐ _b[๐ ^a\stackrel{}{,}๐ ^b]+[๐ _a\stackrel{}{,}๐ _b]๐ซ^a๐ ^b{\displaystyle \frac{2i}{3}}๐ฟ^{abc}๐ _a[๐ _b\stackrel{}{,}๐ _c]\right),`$
$`๐ฎ^{(4)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{๐๐}\left([๐ _a\stackrel{}{,}๐ _b][๐ ^a\stackrel{}{,}๐ ^b]\right).`$
Up to their normalization, the formulas for $`๐ฎ^{(3)}`$ and $`๐ฎ^{(4)}`$ are obvious from eq. (4) and our description of the leading contribution to the 4-point function. The relative normalization of $`๐ฎ^{(3)},๐ฎ^{(4)}`$ and $`๐ฎ_0`$ may be determined most easily by comparison with the abelian case $`๐ฟ_{abc}=0`$. The full effective action
$$๐ฎ=๐ฎ_0+๐ฎ^{(3)}+๐ฎ^{(4)}$$
(4.15)
for the scalar fields on our brane in $`^7\times S^3`$ can be brought into a โgeometricโ form by introducing a field strength
$$๐ฅ_{ab}(๐ )=i๐ซ_a๐ _bi๐ซ_b๐ _a+i[๐ _a\stackrel{}{,}๐ _b]+๐ฟ_{abc}๐ ^c$$
(4.16)
and a non-commutative analogue of the Chern-Simons form
$$\mathrm{๐ข๐ฒ}_{abc}(๐ )=๐ซ_a๐ _b๐ _c+\frac{1}{3}๐ _a[๐ _b\stackrel{}{,}๐ _c]\frac{i}{2}๐ฟ_{abd}๐ ^d๐ _c.$$
(4.17)
Then our effective action can be written as sum of a Yang-Mills theory and of a Chern-Simons theory on the fuzzy sphere,
$$๐ฎ=๐ฎ_{\mathrm{YM}}+๐ฎ_{\mathrm{CS}}=\frac{1}{4}\mathrm{๐๐}\left(๐ฅ_{ab}๐ฅ^{ab}\right)\frac{i}{2}\mathrm{๐๐}\left(๐ฟ^{abc}\mathrm{๐ข๐ฒ}_{abc}\right)$$
(4.18)
In order to render the effective action dimensionless, the expression on the right hand side is to be multiplied with the appropriate power $`(2\pi \alpha ^{})^2`$ of the string tension.
In the computation leading from (4.15) to (4.18) we have added a term of second order in the fields $`๐ _a`$ that vanishes on-shell, i.e. upon using the physical state condition $`๐ซ^a๐ _a=0`$. The main motivation for this modification is that it renders the action off-shell gauge invariant. In fact, some cumbersome but straightforward manipulations show that
$$\delta _\mathrm{\Lambda }๐ฎ_{\mathrm{YM}}(๐ )=\delta _\mathrm{\Lambda }๐ฎ_{\mathrm{CS}}(๐ )=0\text{ where }\delta _\mathrm{\Lambda }๐ _a=i๐ซ_a\mathrm{\Lambda }+i[๐ _a,\mathrm{\Lambda }]$$
for arbitrary $`\mathrm{\Lambda }\text{Mat}(N)\text{Mat}(2\alpha +1)`$. Note, in particular, that the โmass termโ in the Chern-Simons form (4.17) is required by gauge invariance. On the other hand, the effective action (4.18) is the unique combination of $`๐ฎ_{\mathrm{YM}}`$ and $`๐ฎ_{\mathrm{CS}}`$ in which mass terms cancel. As we shall see below, it is this special feature of our action that allows solutions describing translations of the branes on the group manifold. The action $`๐ฎ_{\mathrm{YM}}`$ was already considered in the non-commutative geometry literature , where it was derived from a Connes spectral triple and viewed as describing Maxwell theory on the fuzzy sphere. Arbitrary linear combinations of non-commutative Yang-Mills and Chern-Simons terms were considered in .
There exists a nice way to rewrite our actions such that the statements about gauge invariance become completely obvious. To this end we note that the covariant derivative $`๐ซ_a+[๐ _a,.]`$ can be implemented as a commutator with the element
$$๐ก_a=๐ธ_a+๐ _a\text{ where }๐ธ_a=\frac{1}{\sqrt{2\alpha ^{}}}\mathrm{Y}_a^1,$$
(4.19)
see eq. (3.9). The behavior of $`๐ก_a`$ under gauge transformations is given by $`\delta _\mathrm{\Lambda }B_a=i[B_a,\mathrm{\Lambda }]`$. This property guarantees the gauge invariance of $`๐ฎ_{\mathrm{CS}}`$ and $`๐ฎ_{\mathrm{YM}}`$ once we observe that both actions can be expressed entirely through $`๐ก_a`$ according to
$`๐ฎ_{\mathrm{YM}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{๐๐}\left(๐ฅ_{ab}๐ฅ^{ab}\right)\text{with}๐ฅ_{ab}=i[๐ก_a\stackrel{}{,}๐ก_b]+๐ฟ_{abc}๐ก^c`$ (4.20)
$`๐ฎ_{\mathrm{CS}}`$ $`=`$ $`{\displaystyle \frac{i}{2}}\mathrm{๐๐}\left({\displaystyle \frac{1}{3}}๐ฟ^{abc}๐ก_a๐ก_b๐ก_c{\displaystyle \frac{i}{\alpha ^{}๐}}๐ก_a๐ก^a+{\displaystyle \frac{i}{3\alpha ^{}๐}}๐ธ_a๐ธ^a\right).`$ (4.21)
The two formulas are rather easy to verify by inserting the definition (4.19) of $`๐ก_a`$ and using the algebraic properties of $`๐ธ_a`$. The full low-energy effective action for branes in $`S^3`$ then takes the simple form
$$๐ฎ=\mathrm{๐๐}\left(\frac{1}{4}[๐ก_a\stackrel{}{,}๐ก_b][๐ก^a\stackrel{}{,}๐ก^b]+\frac{i}{3}๐ฟ^{abc}๐ก_a[๐ก_b\stackrel{}{,}๐ก_c]+\frac{1}{6\alpha ^{}๐}๐ธ_a๐ธ^a\right).$$
(4.22)
Up to now, we have only discussed SU(2) branes in bosonic string theory, which are unstable due to the presence of tachyonic modes. In the remainder of this section, we provide the formulas necessary to extend the bosonic discussion from above to the superstring case. As before, we will mostly ignore the curvature of the external manifold $`M^7`$ and pretend that our target has the form $`^7\times S^3`$. In the supersymmetric situation, the SU(2) WZW model is supplemented by three free fermions $`\psi ^a(x)`$, $`a=1,2,3`$, with the following operator product expansions (we use CFT conventions as in eq. (3.5), string units are obtained by rescaling with $`\alpha ^{}`$ in the same way as before)
$$\psi ^a(x_1)\psi ^b(x_2)=\frac{๐}{2}\frac{\delta ^{ab}}{x_1x_2}+\mathrm{},J^a(x_1)\psi ^b(x_2)=\frac{if_c^{ab}}{x_1x_2}\psi ^c(x_2).$$
The world-sheet supercurrent is given by (see e.g. )
$$G=G^7+G_{\mathrm{SWZW}}=G^7+\frac{2}{๐}\delta _{ab}J^a\psi ^b+\frac{4i}{3๐^2}f_{abc}\psi ^a\psi ^b\psi ^c.$$
$`G^7`$ depends on the specific superstring background under consideration. E.g. for the NS 5-brane, we have
$$G^7=\eta _{\mu \nu }J^\mu \psi ^\nu +J^6\psi ^6\frac{2i}{๐}\psi ^6$$
where $`J^\mu =iX^\mu ,\psi ^\mu `$ for $`\mu =0,\mathrm{},5`$ are the free fields from the $`^{5,1}`$ subtheory and where $`J^6=i\varphi `$ is formed from the Feigin-Fuchs boson in the linear dilaton theory, with associated free fermion $`\psi ^6`$. An explicit expression for $`G^7`$ in the AdS$`{}_{3}{}^{}\times T^4`$ model can be found e.g. in ; it is rather similar to the SU(2) supercurrent because of the relation of AdS<sub>3</sub> and the SU(1,1) WZW model.
We are interested in BPS branes here, so in particular the world-sheet supercurrent must be preserved by the corresponding boundary conditions, i.e. we must have $`G(z)=\pm \overline{G}(\overline{z})`$ along the boundary of the world-sheet. It is easy to see that, in the NS 5-brane case, this enforces Neumann conditions on the linear dilaton: All the branes discussed here will end on the stack of NS 5-branes. As usual, the allowed number of Neumann directions in the remaining target dimensions depends on whether we work in IIA or IIB string theory.
It is convenient to pass to a new set of SU(2) currents defined by
$$\stackrel{~}{J}^a:=J^a+\frac{i}{๐}f_{bc}^a\psi ^b\psi ^c,$$
which satisfy the WZW operator product expansion (3.5) at level $`๐2`$ instead of $`๐`$, and which commute with the fermions $`\psi ^a`$. (On AdS<sub>3</sub>, one proceeds analogously for the currents of the SU(1,1) subtheory.) Using the new fields, it becomes clear that the CFTs under consideration split into tensor products of a bosonic part with ten free fermion, and that the same is true for the boundary states describing branes in the NS 5-brane or AdS$`{}_{3}{}^{}\times S^3\times T^4`$ superstring background. As a consequence, space-time supersymmetry charges can be built up from the usual spin fields, and also the behavior under the GSO projection carries over from the flat target case. In particular, the partition functions of our supersymmetric branes in $`M^7\times S^3`$ vanish, meaning that the spectra are tachyon-free and stable. Geometric arguments for stability of branes in $`S^3`$, based on the non-vanishing B-field and on the quantization condition on the brane labels $`\alpha `$, were presented recently in . Our analysis of the equations of motion will, among other things, confirm these results.
Let us now collect formulas for the physical vertex operators necessary to compute the low-energy effective action in the supersymmetric case. Those corresponding to gauge bosons and scalars on the external part of the brane world-volume can be treated as usual, therefore we concentrate on the โHiggs scalarsโ associated with the $`S^3`$-directions. In the picture with superghost number -1, their vertex operators are $`V_H^{(1)}[๐ ](x)=V[e_k]:\psi ^aV[๐ _a]:(x)`$ with corresponding states
$$|V_H^{(1)}[๐ ]=a_a^{Jm}|k\psi _{\frac{1}{2}}^a|Y_m^J.$$
Here, the SU(2) highest weight states are tacitly understood to contain the vacuum of the fermionic Fock space. Moreover, we now approximate the external part $`M^7`$ by flat space-time such that $`k`$ is a $`\stackrel{~}{p}`$-momentum where $`\stackrel{~}{p}`$ is the number of Neumann directions in $`^7`$. As before, the parameters $`๐ _a`$ are subject to the physical state condition $`๐ซ^a๐ _a=0`$.
For the computation of 3- and 4-point functions, we also need vertex operators in the ghost number 0 picture. From the usual commutation relations between $`G_{\mathrm{SWZW}}`$ and $`\psi ^a,J^a`$, one finds the SU(2) component of the corresponding states to be
$`|V_H^{(0)}[๐ ]`$ $`=`$ $`G_{\frac{1}{2}}|V_H^{(1)}[๐ ]`$
$`=`$ $`a_a^{Jm}\left(\stackrel{~}{J}_1^a{\displaystyle \frac{2}{๐}}\delta _{cb}\stackrel{~}{J}_0^c\psi _{\frac{1}{2}}^a\psi _{\frac{1}{2}}^b{\displaystyle \frac{i}{๐}}f_{bc}^a\psi _{\frac{1}{2}}^b\psi _{\frac{1}{2}}^c\right)|Y_m^J.`$
It is now straightforward to compute the 3-point function
$$V_H^{(1)}[๐ _1](x_1)V_H^{(1)}[๐ _2](x_2)V_H^{(0)}[๐ _3](x_3),$$
and the result is given by the terms displayed in eq. (4). In the supersymmetric case, there are no higher order corrections in $`\alpha ^{}`$. Likewise, the leading contribution to the 4-point function is as in the bosonic WZW model.
All in all, the low-energy effective action (4.18), supplemented by the non-abelian Yang-Mills-Higgs theory for the external degrees of freedom, also describes the dynamics of supersymmetric branes in $`^7\times S^3`$. Of course, there are additional terms we did not discuss, which mix the fields from the external gauge theory with the Higgs scalars from the $`S^3`$ factor. To leading order in $`\alpha ^{}`$, they can be copied from the effective action for flat branes in $`^{10}`$. We will not discuss space-time fermions here, whose action could in principle be computed from world-sheet Ramond fields, or by exploiting space-time supersymmetry.
## 5 Classical solutions and bound states
Having found the action that describes the dynamics of branes on SU(2), we can start analyzing the classical equations of motions and their solutions. The equations of motion which derive from the action (4.18) read
$$๐ซ^a๐ฅ_{ab}+[๐ ^a\stackrel{}{,}๐ฅ_{ab}]=0.$$
(5.23)
Geometrically, this means that the curvature of the solutions must be covariantly constant. When we deal with a stack of $`N`$ branes which wrap the fuzzy sphere labeled $`\alpha `$, the equations are to be read as equations in $`\text{Mat}(N)\text{Mat}(2\alpha +1)`$, and $`๐ซ^a=1๐ซ^a`$ acts trivially on the Chan-Paton factors.
We will first concentrate on constant solutions and then extend our discussion to non-constant ones towards the end of the section. On constant fields $`๐ฒ_a`$, i.e. those satisfying $`L^a๐ฒ_b=0`$ for all $`a,b`$, the equations of motion (5.23) simplify to
$$\left[๐ฒ^a\stackrel{}{,}[๐ฒ_a\stackrel{}{,}๐ฒ_b]i๐ฟ_{abc}๐ฒ^c\right]=0.$$
(5.24)
As we will see in a moment, even the constant solutions give rise to a rather non-trivial pattern of D-brane configurations. The equations (5.24) possess two different types of solutions. Obviously, they are solved by any set of commuting matrices $`๐ฒ_a\text{Mat}(N)\mathrm{๐}_{2\alpha +1}`$. These solutions correspond to marginal fields in the SU(2) theory which translate our $`N`$ D-branes in the group target, but they have no influence on the radius. The new positions of the branes are determined by the eigenvalues of the matrices $`๐ฒ_a`$. If some of these eigenvalues agree, smaller stacks of spherical branes appear in the final configuration. All these statements are familiar from D-branes in flat space and they can be verified rigorously along the lines of .
Remarkably, there exists a second type of constant solutions without any flat space analogue. They are given by constant flat connections, or equivalently, by matrices $`๐ฒ_a\text{Mat}(N)\mathrm{๐}_{2\alpha +1}`$ satisfying the commutation relations
$$[๐ฒ_a,๐ฒ_b]=i๐ฟ_{ab}^c๐ฒ_c$$
(5.25)
which, on constant solutions, are equivalent to $`๐ฅ_{ab}(๐ฒ)=0`$. Equations (5.25) simply characterize $`N`$-dimensional representations of su(2). To any such representation $`\pi _N`$ we can assign a partition $`(n_i)_{i=1,\mathrm{},r}`$ of $`N=n_1+\mathrm{}+n_r`$ by choosing $`n_i`$ to be the dimensions of the irreducible sub-representations in $`\pi _N`$. This partition characterizes the original representation uniquely up to equivalence, hence it provides a simple classification for solutions of the second type. Our objective is to identify these solutions with known configurations of branes. Rigorous statements about this identification are difficult since the boundary perturbations generated by the set $`๐ฒ_a`$ are marginally relevant. This implies that there is no continuous trajectory in the moduli space of conformal field theories that we could follow. In this respect, the same problems arise in the study of marginally relevant operators as with open string tachyons and their condensation . Nevertheless, we will provide strong evidence for the pattern of renormalization group fixed points we are about to suggest.
To this end, recall that the state space of our boundary theory is given by $`\text{Mat}(N)_\alpha `$ where $`_\alpha `$ is the state space of a single brane with label $`\alpha `$, and $`\text{Mat}(N)`$ are the wave functions associated with the Chan-Paton factors. The modes $`j_n^a`$ of the chiral fields $`j^a(z)`$ generate an action of $`\widehat{\mathrm{SU}}(2)_๐`$ on the second tensor factor. Moreover, there is an action of su(2) coming with our solution $`๐ฒ_a`$. It is defined through the commutator of $`๐ฒ_a`$ with the Chan-Paton matrices in the first tensor factor $`\text{Mat}(N)`$ of our state space, i.e. $`๐ฒ_a`$ act as $`ad(๐ฒ_a)`$. Once we turn on the perturbation
$$S_{\mathrm{pert}}=๐x:j^aV[๐ฒ_a]:(x)=๐xj^a(x)๐ฒ_a,$$
these two โsymmetriesโ are broken because the perturbing operator couples the SU(2) current to the quantum mechanical su(2) of the Chan-Paton factors. From experience with similar couplings of spin and orbital angular momentum in quantum mechanics, one may suspect that only the sum $`j^a+๐ฒ^a`$ has a good chance to be realized as a symmetry in the perturbed theory. Indeed, the fact that our matrices $`๐ฒ_a`$ obey the equations of motion (5.25) implies that the sum $`j_n^a+ad(๐ฒ^a)`$ gives rise to an action of the current algebra $`\widehat{\mathrm{SU}}(2)_๐`$ with level $`๐`$ on the state space $`\text{Mat}(N)_\alpha `$. Hence, if we assume that the action survives the perturbation, we can use it to determine the representations of $`\widehat{\mathrm{SU}}(2)_๐`$ that are present in the perturbed theory. The decomposition is easy to find with the help of zero modes, and it turns out that there is a unique way of combining their characters back into partition functions of the known D-brane theories. In principle, the partition functions do not quite suffice to recover all information about the D-branes since they are invariant under translations on the background. However, the position of the branes can be detected by bulk fields , which are sensitive only to the scalar part of the Chan-Paton factors. The fact that $`\mathrm{๐๐}๐ฒ_a`$ vanishes for all solutions of eqs. (5.25), shows that, to first order in the perturbative expansion, the branes are not translated, i.e. their classical counterparts are conjugacy classes centered around the origin $`e`$, as listed after eq. (3.4).
The above analysis leads to the following scenario: When we start from a stack of $`N`$ $`\alpha `$-branes centered around $`e`$, the solution $`๐ฒ_a`$ of eqs. (5.25) corresponds to the fixed point of an RG flow that takes us to a new brane configuration consisting of a superposition of spherical branes, again centered around $`e`$. These branes can have different radii, and they can be stacked or just single branes. To be more precise, we use the associated partition $`(n_i)`$ and introduce spins $`J_i=(n_i1)/2`$. Then the boundary state for the final D-brane configuration can be written as sum of โelementaryโ boundary states $`|\beta `$ for single spherical branes,
$$|(๐ฒ_a)=\underset{\beta }{}N_\beta |\beta \mathrm{with}N_\beta =\underset{i}{}N_{\alpha J_i}^\beta .$$
(5.26)
Here, $`N_{\alpha J}^\beta `$ denote the fusion rules of su(2). The multiplicity $`N_\beta `$ is the size of the stack of branes of type $`\beta `$ contained in the final configuration.
As an example, let us consider an initial stack of $`N`$ D0-branes at the origin and choose an irreducible $`N`$-dimensional representation $`๐ฒ_a`$ of su(2). Then, according to our rules, this will decay into a single brane of type $`\beta =(N1)/2`$. In other words, the stack of D0-branes has evolved into a single brane that wraps a non-degenerate 2-dimensional (fuzzy) sphere. One could say that, in this way, we have dynamically created two new dimensions on the expense of breaking the gauge symmetry from U($`N`$) to U(1). Had we used some reducible representation $`๐ฒ_a`$ instead, with partition $`(n_i)_{i=1,\mathrm{},r}`$, stacks of $`N_\beta =\mathrm{\#}\{n_i|n_i=2\beta +1\}`$ branes would have appeared, which wrap the spheres with labels $`\beta `$. In case there were degeneracies among the $`n_i`$, the gauge group would be only partially broken to $`\mathrm{}\mathrm{}_\beta `$U($`N_\beta `$).
We want to use the example of $`N`$ D0-branes blowing up into a single D2-brane to provide more evidence for our identification of the fixed points. The idea is to exploit that we certainly know the tension of all the maximally symmetric branes. If our identification is correct, the leading term in the difference of tensions between the initial and the final brane configuration has to agree with the effective action $`๐ฎ(๐ฒ_c)`$ evaluated at the classical solution. The strategy is similar to the one used in .
The tension of a single brane is essentially given by the logarithm of the so-called $`g`$-factor of the boundary conformal field theory, see e.g. . The latter is defined as the 1-point function of the identity operator , and for a brane of type $`\alpha `$ in the SU(2) WZW model it is given by
$$g_\alpha =\frac{S_{\alpha \mathrm{\hspace{0.17em}0}}}{\sqrt{S_{\mathrm{0\hspace{0.17em}0}}}}==\left(\frac{2}{๐+2}\right)^{\frac{1}{4}}\frac{\mathrm{sin}\frac{(2\alpha +1)\pi }{๐+2}}{\mathrm{sin}^{\frac{1}{2}}\frac{\pi }{๐+2}}.$$
Here $`S_{IJ}`$ denotes the modular S-matrix of the WZW model. For the initial stack of $`N`$ identical D0-branes, the $`g`$-factor is $`g_{N\mathrm{D0}}=Ng_0`$. If we expand the difference of the logarithms in a power series around $`1/๐=0`$ we find to leading order
$$\mathrm{ln}\frac{g_{(N1)/2}}{g_{N\mathrm{D0}}}=\frac{\pi ^2}{6}\frac{N^21}{๐^2}.$$
The minus sign shows that the tension of the final state is lower than the one of the original stack of branes, in agreement with the โ$`g`$-theoremโ stating that the $`g`$-factor should always decrease along the renormalization group trajectory .
Now we have to compare our result on the $`g`$-factors with the value of the action $`๐ฎ(๐ฒ_a)`$ where $`๐ฒ_a`$ is a solution of eqs. (5.25) corresponding to an irreducible $`N`$-dimensional representation of su(2). Because $`๐ฒ_a`$ is constant, all the terms in which $`๐ซ_b`$ acts on $`๐ฒ_a`$ vanish, and after a short computation we obtain
$$๐ฎ(๐ฒ_a)=\frac{(2\pi \alpha ^{})^2}{6\alpha ^{}๐}\mathrm{๐๐}(๐ฒ_c๐ฒ^c).$$
Up to normalization, the term in the argument of the trace is simply the Casimir element of su(2) evaluated in the irreducible representation of spin $`\alpha =(N1)/2`$, i.e. the argument of the trace is proportional to $`\alpha (\alpha +1)=(N+1)(N1)/4`$. Taking the normalization into account, we find
$$๐ฎ(๐ฒ_a)=\mathrm{ln}\frac{g_{(N1)/2}}{g_{N\mathrm{D0}}}$$
as desired. Note that the value of the action (the potential energy of the Higgs fields) at the new fixed point is actually lower than in the original state which means that the $`N`$ D0-branes have formed a bound state.
So far we have only discussed constant solutions to the full equations of motion (5.23). For D0-branes, this discussion is complete since any field $`๐ _a`$ on a single point is constant. On the other hand, we have obtained all other branes $`\alpha `$ as bound states of D0-branes. Combining the two facts, it should be possible to understand non-constant classical solutions, as well. To this end, we first rewrite the equation of motion (5.23) according to
$$\left[๐ก^a\stackrel{}{,}[๐ก_a๐ก_b]i๐ฟ_{abc}๐ก^c\right]=0$$
(5.27)
with $`๐ก_a=๐ธ_a+๐ _a`$ as before. Now the general equations of motion (5.23) have become formally identical to (5.24) which govern the dynamics of D0-branes, and solving them amounts to finding matrices $`๐ก_a\text{Mat}(N)\text{Mat}(2\alpha +1)`$ that satisfy our familiar eqs. (5.27). From any such solution $`๐ก_a`$ we can then obtain a field $`๐ _a=๐ก_a๐ธ_a`$ that will carry us into a new RG fixed point upon perturbation. Hence, it only remains to identify this fixed point. In case that $`๐ก_a`$ is a representation of su(2), we claim that the fixed point is determined by the partition $`(n_i)`$ characterizing the representation $`๐ก_a`$ of su(2) along with the rules we have used for stacks of D0 branes above. This means that there will appear $`N_\beta =\mathrm{\#}\{n_i|n_i=2\beta +1\}`$ branes wrapping the spheres with labels $`\beta `$ in the final configuration.
Note that this pattern is consistent with the one proposed in (5.26). In fact, if $`๐ _a=๐ฒ_a`$ is constant, i.e. if it commutes with $`๐ธ_a`$, then the elements $`๐ก_a`$ generate the tensor product of the $`(2\alpha +1)`$-dimensional irreducible representation with the representation $`๐ฒ_a`$. Hence, the partition associated with $`๐ก_a`$ can be obtained from the one that comes with $`๐ _a`$ upon fusing with the spin $`\alpha `$ representation.
Our claim on the nature of the fixed points described by su(2) representations $`๐ก_a`$ can be substantiated with the same methods as before. In particular, one can again compare the $`g`$-factors of the proposed fixed points with the value of the effective action. When evaluated on a solution of our equation of motion (5.27), the latter becomes
$$๐ฎ(๐ก_a)=\frac{1}{12}\mathrm{๐๐}\left([๐ก_a๐ก_b][๐ก^a๐ก^b]+\frac{2}{\alpha ^{}๐}๐ธ_a๐ธ^a\right).$$
(5.28)
The formula shows that commuting matrices $`๐ก_a`$ cause the action to increase, i.e. they lead to a final configuration which has a larger mass than the one we started with. On the other hand, the lowest action among all the solutions we have described is assumed for $`๐ก_a`$ being an irreducible representation of su(2). This corresponds to the stack of branes merging into a single brane with $`\beta =N\alpha +(N1)/2`$.
In addition, one may compute the second variation of the action and evaluate it at an $`N(2\alpha +1)`$-dimensional su(2) representation $`๐ก_a`$ to get
$$\delta ^2๐ฎ=\mathrm{๐๐}\left([๐ก_a\stackrel{}{,}\delta ๐ก_b][๐ก^a\stackrel{}{,}\delta ๐ก^b]+[๐ก_a\stackrel{}{,}\delta ๐ก^a][๐ก_b\stackrel{}{,}\delta ๐ก^b]\right).$$
Using the fact that the map $`๐ [๐ก_a,[๐ก^a,๐ ]]`$ is invertible on all traceless matrices $`๐ `$, one can easily see that the su(2) representations $`๐ก_a`$ represent a local minimum of the action (except from the obvious zero-modes associated with rigid translations), in agreement with the results of on the stability of the spherical branes in WZW models.
The previous analysis of non-constant solutions may be turned around and used to โderiveโ the equations of motion (5.23) for an arbitrary brane of type $`\alpha `$ from those of D0-branes. <sup>1</sup><sup>1</sup>1This argument was suggested to us by J. Maldacena. In fact, we may think of the fields $`๐ _a`$ on the $`\alpha `$-brane as fluctuations around the constant solution $`๐ฒ_a`$ which takes us from $`N`$ D0-branes to the single D2-brane. Now we recall that $`๐ฒ_a`$ is an irreducible $`(2\alpha +1)`$-dimensional representation of su(2). The world-volume algebra of the $`\alpha `$-brane, on the other hand, contains the elements $`๐ธ_a`$ which also generate an irreducible su(2) representation of dimension $`2\alpha +1`$. Since any two irreducible su(2) representation of the same dimension are equivalent, we can identify $`๐ฒ_a`$ with $`๐ธ_a`$ so that the idea of studying small fluctuations around $`๐ฒ_a`$ motivates to insert the combination $`๐ธ_a+๐ _a`$ into the equations (5.24). The resulting equations are given by (5.27) and do indeed describe the full dynamics of the $`\alpha `$-brane.
Let us finally note that our discussion of the constant solutions coincides with the โabsorption of boundary spinโ that has appeared in the work of Affleck and Ludwig on the multichannel Kondo-problem . In that context, an irreducible representation $`๐ฒ_a\text{Mat}(N)`$ is reinterpreted as the quantum mechanical spin variable of a spin $`J=(N1)/2`$ impurity. According to Affleck and Ludwig, this system evolves into a new low-temperature fixed point in which the boundary spin gets absorbed.
## 6 Conclusions and outlook
In this paper we have studied the dynamics of branes in $`M^7\times S^3`$ which, classically, wrap some $`S^2S^3`$, focussing on the limit where the radius $`R\sqrt{\alpha ^{}๐}`$ of $`S^3`$ grows very large. The fact that the spherical conjugacy classes become fuzzy spheres in the quantum theory has a number of consequences for the low-energy effective action:
First, the gauge fields on the external part of the branes become non-abelian already on a single brane. The size of the gauge group U$`(N(2\alpha +1))`$ depends on the number $`N`$ of coinciding branes and on the label $`\alpha `$ of the fuzzy sphere that is wrapped in $`S^3`$.
Secondly, the Higgs potential for the scalars associated with the internal $`S^3`$ consists of a non-commutative Yang-Mills theory on the fuzzy sphere plus a non-commutative analogue of the Chern-Simons action. As expected on general grounds, this LEEA enjoys invariance under non-commutative gauge transformations.
We have studied solutions to the equations of motion derived from the non-commutative part of the effective action. We found that a certain class of constant field configurations, namely those obtained from pairwise commuting constant gauge fields, are associated with moduli that trigger rigid translations of branes in the $`S^3`$ background. The more interesting type of solutions is given by reducible or irreducible representations of su(2) and has no analogue for flat branes in $`^d`$. These solutions describe fixed points of RG flows driven by marginally relevant perturbations of the world-sheet theory. At the fixed points, the initial stack of identical branes has evolved into a bound state of SU(2) branes, whose precise form is determined by the su(2) representation. In contrast to the โconstant commuting caseโ, the new type of solutions given in (5.25) can change the radii of the branes. In this way, we can in particular interpret 2-dimensional spherical branes as bound states formed from a stack of D0-branes.
As we pointed out above, gauge theories on fuzzy spheres were discussed in the literature before , and the field equations that were derived from such actions are very similar to the equations we have studied here. The non-commuting solutions (5.25) are a common feature of these cases, but the model we have derived here from perturbative string theory is distinguished by the property that it also admits the commuting solutions corresponding to rigid movements in $`S^3`$.
Based on what we have described above, one cannot exclude the existence of additional solutions which could lead the system into new types of bound states. Even if such configurations exist, it is likely that the single spherical brane realizes the global minimum of the action. We have not been able to prove this in the general case but at least it is true for $`N(2\alpha +1)=2`$.
Extending the present analysis to finite radius $`R\sqrt{\alpha ^{}๐}`$ is another challenge. From the results of one may be tempted to believe that part of the picture could survive if the Lie algebra su(2) is replaced by the corresponding quantum group and its representation theory. The relation with quantum group representations was also stressed in ; their importance also transpires from the general investigations in . Note that in the Kondo problem, the level $`๐`$ counts the number of conduction bands. Hence, Kondo physics explores the region of small level. In this context one may recall that for $`N=๐n/2`$ and $`n`$ integer, which of course requires to go beyond a perturbative analysis in $`1/๐`$, the known fixed-points are described by free rather than interacting theories . This suggests that new phenomena occur for stacks of D0-branes on a sphere of finite size. For instance, the analysis of Affleck and Ludwig shows that a stack of $`N`$ D0-branes can evolve into a single spherical brane with label $`\alpha `$ where $`\alpha ๐/2`$ is determined from $`N`$ by the rule $`N=๐n/2+\alpha `$. In view of the work , which compares โgiant gravitonsโ with spherical branes on SU(2), such finite level effects should be related to the โstringy exclusion principleโ.
As we have pointed out before, most of our statements admit obvious generalizations to other group manifolds. The necessary CFT ingredients can be recovered e.g. from . In view of the perturbative treatment of and of the general definition of non-commutative world-volumes from boundary CFT given in , we believe that the lessons we have learned are much more universal. Flows from stacks of D0-branes to branes whose classical counterparts have higher-dimensional world-volume should be a common phenomenon in string compactifications. It is a distinguished feature of string theory that the world-sheet perspective provides natural descriptions of effects that appear like a โtopology changeโ in the space-time picture, e.g. the dynamical generation of extra world-volume dimensions as discussed above, or the rotation of Dirichlet into Neumann conditions for free bosons by marginal operators . We hope that extensions of our results may also provide some quantitative insight into tachyon condensation and bound state formation on other more complicated string backgrounds which possess an exact conformal field theory description. This applies especially to Gepner models and their associated D-brane theories . Another step in this direction was taken recently in , and we hope to return to these issues in the future.
###### Acknowledgments.
We would like to thank I. Brunner, M. Douglas, S. Fredenhagen, J. Frรถhlich, K. Gawศฉdzki, H. Grosse, J. Louis, D. Lรผst, J. Madore, J. Maldacena, G. Reiter, S.-J. Rey, M. Rozali, J. Teschner and S. Theisen for useful and stimulating discussions. This work was started during a stay of the three authors at the ESI for Mathematical Physics in Vienna. |
warning/0003/hep-ex0003028.html | ar5iv | text | # References
DFUB 99/20
SLIM 99/1
SEARCH FOR โLIGHTโ MAGNETIC MONOPOLES
The SLIM Collaboration
D. Bakari<sup>1,4</sup>, S. Cecchini<sup>1,a</sup>, H. Dekhissi<sup>1,4</sup>, J. Derkaoui<sup>1,4</sup>, G. Giacomelli<sup>1</sup>, M. Giorgini<sup>1</sup>, J. McDonald<sup>3</sup>, G. Mandrioli<sup>1</sup>, S. Manzoor<sup>1,5</sup>, A. Margiotta<sup>1</sup>, A. Marzari-Chiesa<sup>2</sup>, J. Nogales<sup>7</sup>, M. Ouchrif <sup>1,4</sup>, L. Patrizii<sup>1</sup>, J. Pinfold<sup>3</sup>, V. Popa<sup>1,6</sup>, O. Saavedra<sup>2</sup>, P. Serra<sup>1</sup>, M. Spurio<sup>1</sup>, R. Ticona<sup>7</sup>, V. Togo<sup>1</sup>, A. Velarde<sup>7</sup>, E. Vilela<sup>1</sup> and A. Zanini<sup>2</sup>
1. Dipar.to di Fisica dellโUniversitร di Bologna and INFN, 40127 Bologna, Italy
2. Dipar.ti di Fisica Sperimentale e Generale dellโUniversitร di Torino and INFN, 10125 Torino, Italy
3. Centre for Subatomic Research, Univ. of Alberta, Edmonton, Alberta T6G 2N4, Canada
4. Faculty of Sciences, University Mohamed I, B.P. 424, Oujda, Morocco
5. RPD, PINSTECH, P.O. Nilore, Islamabad, Pakistan
6. Institute for Space Sciences, 76900 Bucharest, Romania
7. Laboratorio de Fisica Cosmica de Chacaltaya,UMSA, La Paz, Bolivia
$`a`$ Also Istituto TESRE/CNR, 40129 Bologna, Italy
## Abstract
We propose to implement one passive nuclear track detector array of $`400`$ m<sup>2</sup> at the Chacaltaya High Altitude Laboratory (5230 m a.s.l.). The main purposes of the experiment concern the searches for magnetic monopoles of relatively low masses at the Parker bound level, searches for low mass nuclearites, and searches for some supersymmetric dark matter candidates (Q-balls).
1. Introduction
The search for magnetic monopoles (MMs) in the penetrating cosmic radiation remains one of the main items of non-accelerator particle astrophysics.
Grand Unified Theories (GUT) of electroweak and strong interactions predict the existence of superheavy magnetic monopoles with masses larger than $`10^{16}`$ GeV . They would have been produced at the end of the GUT epoch, at a mass scale $`10^{14}`$ GeV and the cosmic time of $`10^{34}`$ s. Such monopoles cannot be produced with existing accelerators, nor with any foreseen for the future. The MACRO experiment is well suited for their study and is providing the best experimental limits .
Lower mass monopoles, proposed by many authors, require a phase transition in the early universe in which a semisimple gauge group yields a U(1) factor at a lower energy scale . MMs with masses around $`10^6รท10^{10}`$ GeV have been proposed \[3-6\]. A MM is a topological point defect; an undesirable large number of relatively light monopoles may be gotten rid of by means of higher dimensional topological defects (strings, walls) .
One of the recent interests in relatively low mass MMs is connected also with the possibility that relativistic MMs could be the source of the highest energy cosmic rays, with energies larger than $`10^{20}`$ eV \[3-5\]. For monopoles one knows possible acceleration mechanisms: since the basic magnetic charge should be very large, relatively light monopoles can be accelerated to relativistic velocities and to energies of the order of $`10^{20}`$ GeV in one coherent domain of the galactic magnetic field, or in the intergalactic field, or in many astrophysical sites, like in the magnetic field of Active Galactic Nuclei (AGN) and even of neutron stars. The next problem is how these monopoles can interact in the upper atmosphere and yield electromagnetic showers. This is for istance possible if a monopole forms a bound state with a proton (a dyonic system) which then may interact with a cross section typical of a relativistic hadron ($`\sigma 10^{26}`$ cm<sup>2</sup>) \[3-5\]. Monopole masses of $`10^6รท10^{10}`$ GeV could be consistent with a flux at the Parker limit .
We must also remember the possibility that MMs could be multiply charged, $`g=2g_D`$, as in some SUSY theories, and $`g=3g_D`$, as in some superstring models ($`g_D=\mathrm{}c/2e=68.5e`$ is the basic Dirac monopole charge) and that the basic charge could be $`1/3e`$ .
We propose a search for relatively light MMs with an array of 400 m<sup>2</sup> of passive nuclear track detectors deployed at the high altitude Chacaltaya lab (5230 m high above sea level). In $`>4`$ years of operation we should be able to reach a sensitivity at the level of the Parker bound, i.e. $`10^{15}`$ cm<sup>-2</sup> s<sup>-1</sup> sr<sup>-1</sup>. Byproducts of this MM search are the searches for relatively light nuclearites and Q-balls .
We recall that nuclearites (strangelets, strange quark matter) are nuggets of strange quark matter (aggregates of $`u`$, $`d`$, and $`s`$ quarks in equal proportions); they could be the ground state of QCD and could be part of the cold dark matter, and could have typical galactic velocities $`\beta 10^3`$ .
Q-balls are supersymmetric coherent states of squarks, sleptons and Higgs fields, predicted by minimal supersymmetric generalizations of the Standard Model; they could be copiously produced in the early universe. Relic Q-balls are also candidates for the cold dark matter .
Since both nuclearites and charged Q-balls lose a large amount of energy for $`\beta >4\times 10^5`$ they would be easily detectable with the proposed track-etch system.
An exposure at a high altitude laboratory would allow to search for MMs of lower masses, higher magnetic charges and lower velocities , see Figs. 1 and 2. The same holds for lighter nuclearites and Q-balls. For low mass nuclearites one would reach a level of sensitivity more than one order of magnitude lower than any of the existing limits .
Fig. 3 shows the accessible region in the plane (mass, $`\beta `$) for nuclearites, at MACRO depth, at ground level (under 1000 g cm<sup>-2</sup> of atmosphere), at the Chacaltaya altitude (540 g cm<sup>-2</sup> of atmosphere) and at 20 km height assuming that the nuclearites have standard energy losses . Lower mass nuclearites should be much more abundant than higher mass ones .
The high altitude exposure would allow detection of the above mentioned particles even if they had strong interaction cross sections which could prevent them from reaching the earth surface. From this point of view, it is important that the site be at the highest altitude. (On this point we have asked the opinions of several colleagues. In particular S. Glashow, J. Steinberger and A. De Rรบjula adviced us to search for the above mentioned objects at as high an altitude as possible.)
Experimental data obtained at the highest altitude laboratories suggest the existence of โCentauro eventsโ and other exotic events. It has also been suggested that nuclearites with mass number of only few hundred could have nuclearite - air nuclei collisions at high altitudes in which the baryonic number of the nuclearite reduces by about the mass number of the target nucleus; this effect can be neglected for large mass nuclearites but it could seriously alter the energy loss of nuclearites with $`A1000`$ . The nuclearites would decrease in $`A`$ in successive interactions with air nuclei until reaching some critical value, $`A_{crit}320`$, below which they disintegrate into nucleons. This mechanism could be tested at the highest altitude stations.
According to ref. the initial mass number that a nuclearite should have at the entry in the atmosphere in order to reach the Chacaltaya altitude (540 g cm<sup>-2</sup>) before disintegrating into normal baryons is $`A3000`$, while to reach 450 g cm<sup>-2</sup> altitude the initial mass should be about $`A1700`$. Assuming that the abundance of nuclearites outside the atmosphere has the same $`A`$ dependence as the abundance of elements in the Universe, the flux for $`A1700`$ would be about 100 times larger than the flux of nuclearites with $`A3000`$ .
We hope to instrument a small area ($``$ 2 m<sup>2</sup>) with many layers of thin CR39 and other detectors; this also applies to a very small detector at higher altitudes.
2. Experimental method
In order to reach the goal of a sensitivity at the level of the Parker bound , one needs a surface detector of the least 400 m<sup>2</sup> and operation for at least four years. This detector also yields good limits on lighter nuclearites and Q-balls.
In order to achieve the best redundancy and โconvincingnessโ the best detector should have redundant types of subdetectors, such as those presently used by the MACRO experiment at Gran Sasso: i) liquid scintillators for wave form shape and time-of flight (ToF) informations, ii) tracking system to ensure single space track and single โtime trackโ (for slow monopoles); iii) passive nuclear track detectors for space track and restricted energy loss analyses . This solution would also yield byproducts on cosmic ray physics , but it would be expensive and complicated and it is not needed for the limited purposes of this proposal.
For the near future the simplest possibility is the use of several layers of different passive nuclear track detectors.
An exposure at a high altitude would effectively lower the monopole mass threshold and it would offer the new possibility of searching for any heavily ionizing object present in the cosmic radiation and which has a strong interaction cross section. This includes light MMs with attached $`p`$ or nuclei, dyons, nuclearites and Q-balls; in fact one might consider this last point as one of the main reasons for such a search at high altitude.
The CR39 nuclear track detector allows to search for magnetic monopoles with one unit Dirac charge ($`g_D`$), for $`\beta =v/c`$ around $`10^4`$ and for $`\beta >10^3`$, the whole $`\beta `$-range of $`4\times 10^5<\beta <1`$ for MMs with $`g2g_D`$, for dyons, for nuclearites and for Q-balls.
We are presently making tests by exposing nuclear track detectors in Bologna and at the Chacaltaya mountain station, in order to study the effects of possible backgrounds and of possible climatic conditions.
3. Proposal
As already stated, we would like to implement passive nuclear track detectors of 400 m<sup>2</sup> at mountain altitude. The track-etch detectors could be organised in modules of 24 cm $`\times `$ 24 cm, each made of 3 layers of CR39, 3 layers of polycarbonate and of an aluminium absorber 1 mm thick; this module would be placed in an aluminized polyethilene bag filled with dry air. These bags reduce by about one order of magnitude the radon background. The CR39 is the main nuclear track detector; the polycarbonate has a higher threshold, and it is useful for high velocity monopoles and for nuclearites and Q-balls with $`\beta >10^4`$.
The best and least expensive CR39 is produced by the Intercast Europe Co. of Parma. We are in a position to obtain very good material, controlled continuously by us, and at the best possible price.
A program for the analysis of various types of polycarbonate (Lexan, Makrofol) is well under way. So far the best results were obtained with Makrofol (made by Bayer) of 0.5 mm thickness. There is no problem for the availability and cost either of Lexan or Makrofol.
Once exposed, CR39 and the Makrofol should be etched. Presently there are two etching facilities, one at Gran Sasso (the apparatus was built in Torino) and one in Bologna. The etching capacity of the Gran Sasso apparatus is about 26 m<sup>2</sup>/month for โstrongโ etching of the first layer of CR39. The Bologna apparatus is presently used for โnormalโ etching of the second CR39 layers; we are planning to use it also for strong etching to complement the Gran Sasso apparatus. The global etching rate would be 42 m<sup>2</sup>/month At the quoted rate, we should be able to complete the etching of the MACRO nuclear track detector in about 2 years.
For the SLIM experiment we plan to etch small samples in the next few years, and start the main etching after the completion of the MACRO effort. We propose to etch about 150 m<sup>2</sup>/year by using only the Bologna apparatus. The collaborators of the Pinstech Lab. could take care of the exposure, etching and analysis of about 100 m<sup>2</sup> of detector.
For the etching and for the analysis of the detector we may use the same methods presently used for MACRO; we shall also study other more authomated methods.
After exposure, the first sheet of each CR39 module is etched using a โstrong etchingโ at 80C in an 8N water solution of NaOH.
The etched CR39 is analyzed, first quickly using a simple method with a light source and a large lens and observing the CR39 foil in transparency; a more accurate analysis is later performed with a large field binocular microscope. At present, for the MACRO CR39, which has an average exposure time of 7 years, we find โcandidatesโ in about 8% of the sheets. In these cases we etch the second CR39 foil of the interested module at 70C in a 6N NaOH water solution (this is presently done in Bologna).
We need to make regular small exposures of CR39 and Makrofol to relativistic heavy ions in order to check the quality of the material used, its stability in time, the absence of fading effects, etc. . One main calibration test using Fe ions of 1 GeV/nucleon from the Brookhaven AGS is in progress.
5. The collaboration
The present collaboration involves groups from INFN and the Universities of Bologna and Torino in Italy, the University of Alberta in Canada, the University of Oujda in Morocco, the Pinstech Lab. in Islamabad, Pakistan, the Institute for Space Sciences of Bucharest, Romania, and the Laboratorio de Fisica Cosmica de Chacaltaya in La Paz, Bolivia. The main contributions from the Bucharest and Oujda groups are with personnel often stationed in Bologna in the context of bilateral agreements.
7. Conclusions
We propose to install a nuclear track detector of 400 m<sup>2</sup> at the Chacaltaya high altitude lab. The detectors will be operated for at least 4 years. They will allow for a search for light monopoles and dyons at the level of the Parker bound, that is at a flux of a $`10^{15}`$ cm<sup>-2</sup> s<sup>-1</sup> sr<sup>-1</sup>. A similar level of sensitivity will be obtained for relatively light nuclearites and for Q-balls.
Acknowledgements. We would like to acknowledge fruitful discussions with A. De Rรบjula, S. Glashow, J. Steinberger, and other colleagues.
We thank the University of Bologna, the ICTP-TRIL program, Worldlab and FAI-INFN for providing fellowships and grants for non Italian scientists. |
warning/0003/cond-mat0003430.html | ar5iv | text | # Pseudospin Anisotropy Classification of Quantum Hall Ferromagnets
## I Introduction
Studies of magnetic phenomena in semiconductors have opened fruitful new ways to explore the subtleties of quantum magnetism. In quantum Hall samples, the tunability of semiconductor electronic systems and the quantization of single-particle energies into macroscopically degenerate Landau levels (LLs) combine to open up a rich and varied phenomenology. The effective zero width of electronic energy bands enhances the role of inter-particle interactions and can frequently lead to the formation of ordered many-particle ground states, including ferromagnetic ones.
Most studies of quantum Hall ferromagnets (QHFs ) have focused on spontaneous spin-alignment in a single-layer two-dimensional (2D) electron system at LL filling factor $`\nu =1`$. In this case it turns out that electron-electron interactions favor fully aligned electron spins even in the limit of vanishingly small Zeeman coupling and the ferromagnetic ground state of the system is described exactly by Hartree-Fock (HF) theory. Because of the near spin-independence of the Coulomb interaction, the $`\nu =1`$ single layer QHF is a Heisenberg-like isotropic two-dimensional ferromagnet. One of the unique properties of this simple itinerant electron ferromagnet is that its instantons (Skyrmions) carry charge and can be observed in the ground states at filling factors slightly deviating from 1.
The notion of the QHF can be generalized however. It turns out that, at least according to HF theory, broken symmetry ground states occur at integer filling factors in quantum Hall systems any time two or more valence LLs are degenerate and the number of electrons is sufficient to fill only some of the LLs. In effect, electrons in the ordered state occupy spontaneously generated LLs that are linear combinations of the single-particle levels chosen to minimize the electron interaction energy.
The simplest example of a pseudospin QHF obtains at $`\nu =1`$ in balanced bilayer 2D systems where the single-particle LLs in the two-layers are degenerate. In the ordered ground state the electrons occupy a LL which is a linear combination of the isolated layer levels, forming a state with spontaneous inter-layer phase coherence. Recently, Josephson-like behavior seen in 2D-to-2D tunneling spectroscopy studies of bilayer systems has provided a direct manifestation of collective behavior generated by this broken symmetry. In $`\nu =1`$ bilayer systems, the broken symmetry state minimizes the Hartree energy cost by distributing charge equally between layers and gives up part of the intra-layer exchange energy while gaining more in inter-layer exchange energy. Unlike the single layer $`\nu =1`$ QHF, the ordered HF ground state is not exact in this case, and the order predicted by HF theory can be destroyed. With decreasing inter-layer exchange energy quantum fluctuations around the mean-field ordered state become more important and for layer separations larger than approximately two magnetic lengths fluctuations destroy the spontaneous coherence. The corresponding order-disorder quantum phase transition has been observed experimentally. In $`\nu =1`$ bilayer QHFs, it is the layer degree of freedom which is represented as a pseudospin-1/2. With this mapping the phase coherent state is equivalent to a spin-1/2 easy-plane ferromagnet. Finite temperature Kosterlitz-Thouless phase transition, continuous quantum phase transition induced by in-plane magnetic field, and macroscopic collective transport effects are are among the remarkable phenomena which have been studied on bilayer QHFs.
Recent experiments in single layer and bilayer 2D systems at even-integer filling factors have further enlarged the field of quantum Hall ferromagnetism. It has been shown that easy-axis ferromagnetic ground states can occur at higher filling factors when LLs with different orbit radius quantum numbers are brought close to alignment. In HF language, easy-axis anisotropy means that many-body states with either of the two aligned LLs completely filled and the other empty is energetically more favorable than the coherent superposition state. The easy-axis pseudospin anisotropy occurs in this case because intra LL exchange is stronger than exchange between particles from LLs with different orbit radius quantum numbers. Transport measurements have demonstrated that easy-axis QHFs exhibit hysteresis with a complicated phenomenology, presumably associated with an interplay between disorder and domain-morphology similar to that in conventional thin film magnets.
In this paper we classify QHFs according to their pseudospin anisotropy energies as either isotropic, easy-axis, or easy-plane systems. We report on a HF based analysis which predicts how the class of broken symmetry ground state depends on the nature of the crossing LLs. In some cases, competing effects allow the anisotropy energy to be tuned continuously generating a zero-temperature quantum phase transition between different classes of states. We consider only cases where no more than two LLs are nearly degenerate and the number of electrons is sufficient to occupy one of them. We will always assume that lower energy LLs, if present, are completely full and higher energy LLs are completely empty; coupling to these remote LLs can usually be treated perturbatively if necessary, although we do not do so explicitly here. An important example of an instance in which more than two LLs are close to degeneracy pertains in double-layer systems with weak tunneling and weak Zeeman coupling; we do not treat this or other more complex cases with many degenerate LLs in this paper. In Section II we precisely define the pseudospin language we use in which one of the LLs is referred to as the pseudospin-up state and the other LL as the pseudospin-down state. Since we assume that the magnetic field is perpendicular to the 2D electron layer, growth direction and in-plane degrees of freedom decouple. The pseudospin quantum number then subsumes real-spin, orbit radius, and growth direction (subband) degrees of freedom. To make the discussion more transparent we concentrate on a system consisting of two nearby infinitely narrow 2D layers, the simplest model which has a non-trivial growth direction degree of freedom. Comments are made throughout the text about realistic samples with more complicated geometries. In Section III we derive a general expression for the HF ground state energy in the pseudospin ferromagnetic state. Section IV summarizes the rather cumbersome evaluation of Coulomb interaction matrix elements. Readers not interested in technical details of the calculation are encouraged to skip to Section V where we present our conclusions concerning pseudospin magnetic anisotropy of single layer and bilayer QHFs. This section includes phase diagrams which show the regimes of physically tunable parameters with easy-axis and easy-plane anisotropies. Symmetry breaking fields and the dynamics of pseudospin reversal are discussed in Section VI. Finally, we conclude in Section VII with a brief summary of the main results of our paper.
## II Pseudospin representation
The pseudospin language was introduced to the description of broken symmetry states in the quantum Hall regime, in order to draw on the analogy between double quantum well systems at $`\nu =1`$ and 2D ferromagnets. In this work the pseudospin degree of freedom represented the layer index of a bilayer system. Here we allow the pseudospin index to have a more general meaning. To establish terminology, it is useful to recall the single-particle spectrum of a bilayer 2D system subject to a perpendicular magnetic field. Quantum well subbands of individual layers can be mixed by interlayer tunneling and shifted by the application of a bias potential. We limit our attention to the usual case where only the lowest electric subband of either quantum well is occupied and for explicit calculations use a zero-width quantum well model. Allowing for external bias and for tunneling between the wells (see Fig. 1), the bilayer subband wavefunctions of the zero-width model are
$$\lambda _{\pm 1}(z)=\frac{1}{\sqrt{2}}\left[(1r_\mathrm{\Delta })^{1/2}\delta (z)\pm (1\pm r_\mathrm{\Delta })^{1/2}\delta (zd)\right],$$
(1)
where $`r_\mathrm{\Delta }=\mathrm{\Delta }_V/(\mathrm{\Delta }_V^2+\mathrm{\Delta }_t^2)^{1/2}`$, $`\mathrm{\Delta }_V`$ is the bias potential, $`\mathrm{\Delta }_t`$ the tunneling gap at zero bias, and $`d`$ is the layer separation. To account for specific experimental samples, finite width effects can be incorporated by replacing the wavefunctions (1) with electric subband wavefunctions calculated using the self-consistent local-spin-density-approximation (LSDA) model.
In the Landau gauge the wavefunctions in the 2D plane take a form $`\varphi _{n,s,k}(x)\mathrm{exp}(iky)/\sqrt{L_y}`$, where
$`\varphi _{n,s,k}(x)`$ $`=`$ $`\left[\pi \mathrm{}^22^{2n}(n!)^2\right]^{1/4}H_n\left({\displaystyle \frac{x\mathrm{}^2k}{\mathrm{}}}\right)`$ (2)
$`\times `$ $`\mathrm{exp}\left[{\displaystyle \frac{(x\mathrm{}^2k)^2}{2\mathrm{}^2}}\right],`$ (3)
$`k`$ is the wavevector label which distinguishes states within a LL, $`n=0,1,\mathrm{}`$ is the orbit radius quantum number, $`\mathrm{}`$ is the magnetic length, and we have also explicitly included the real-spin, $`s=\pm 1/2`$, degree of freedom. The single-particle energy spectrum consists of discrete LLs
$$E_{\xi ,n,s}=\frac{\xi }{2}(\mathrm{\Delta }_V^2+\mathrm{\Delta }_t^2)^{1/2}+\mathrm{}\omega _c(n+\frac{1}{2})s|g|\mu _BB,$$
(4)
where $`\omega _c`$ is the cyclotron frequency and the last term is the real-spin Zeeman coupling. Each LL has a macroscopic degeneracy with the number of orbital states per level $`N_\varphi =AB/\mathrm{\Phi }_0`$, where $`A`$ is the system area, $`B`$ is the field strength, and $`\mathrm{\Phi }_0`$ is the magnetic flux quantum.
The many-body broken symmetry states we study in the following sections occur when two LLs are brought close to alignment while remaining sufficiently separated from other LLs. In our calculations, each LL can have one of two possible subband indices, one of two possible spin indices, and any value for the 2D cyclotron orbit kinetic energy index. The two crossing LLs can differ in any or all of these labels. We label one of the two levels as the pseudospin-up ($`\sigma =`$) state and the other level as the pseudospin down ($`\sigma =`$) state. We truncate the single-particle Hilbert space by ignoring higher LLs and introducing effective one-body fields that account for the effect of electrons in lower LLs on the two pseudospin states. Within this model the set of single-particle states reduces to following wavefunctions
$$\psi _{\sigma ,k}(\stackrel{}{r})=\lambda _{\xi (\sigma )}(z)\varphi _{n(\sigma ),s(\sigma ),k}(x)\frac{\mathrm{exp}(iky)}{\sqrt{L_y}}.$$
(5)
A particle with the pseudospin oriented along a general unit vector $`\widehat{m}=(\mathrm{sin}\theta \mathrm{cos}\phi ,\mathrm{sin}\theta \mathrm{sin}\phi ,\mathrm{cos}\theta )`$ is described by
$$\psi _{\widehat{m},k}(\stackrel{}{r})=\mathrm{cos}\left(\frac{\theta }{2}\right)\psi _{,k}(\stackrel{}{r})+\mathrm{sin}\left(\frac{\theta }{2}\right)e^{i\phi }\psi _{,k}(\stackrel{}{r}).$$
(6)
## III Many-body Hamiltonian and HF total energy
In the HF approximation, the QHF has a single Slater determinant state with the same pseudospin orientation for every orbital $`k`$. In this section we derive general expressions for the dependence of the many-electron state energy on pseudospin orientation. It is convenient to express the many-body Hamiltonian using Pauli spin-matrices $`\tau _x`$, $`\tau _y`$, and $`\tau _z`$ and the $`2\times 2`$ identity matrix which we label $`\tau _\mathrm{๐}`$. In this representation the Hamiltonian reads
$`H`$ $`=`$ $`{\displaystyle \underset{i=\mathrm{๐},x,y,z}{}}{\displaystyle \underset{k=1}{\overset{N_\varphi }{}}}{\displaystyle \underset{\alpha ,\alpha ^{}=1}{\overset{2}{}}}b_i\tau _i^{\alpha ^{},\alpha }c_{\sigma (\alpha ^{}),k}^{}c_{\sigma (\alpha ),k}`$ (7)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i,j=\mathrm{๐},x,y,z}{}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{k_1,k_1^{},}{k_2,k_2^{}=1}}{\overset{N_\varphi }{}}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{\alpha _1,\alpha _1^{},}{\alpha _2,\alpha _2^{}=1}}{\overset{2}{}}}W_{i,j}^{k_1^{},k_2^{},k_1,k_2}\tau _i^{\alpha _1^{},\alpha _1}\tau _j^{\alpha _2^{},\alpha _2}`$ (8)
$`\times `$ $`c_{\sigma (\alpha _1^{}),k_1^{}}^{}c_{\sigma (\alpha _2^{}),k_2^{}}^{}c_{\sigma (\alpha _2),k_2}c_{\sigma (\alpha _1),k_1},`$ (9)
where $`\sigma (1)=`$ and $`\sigma (2)=`$. The one-body terms $`b_i`$ include, in general, the external bias potential, tunneling, cyclotron and Zeeman energies and also the mean-fields from interactions with electrons in the frozen LLs lying below the $`\sigma =`$ and $``$ levels. We will give an explicit expression for $`b_i`$ in section VI. Here and in the following two sections, we concentrate on the two-body terms in the Hamiltonian (9).
The potentials $`W_{i,j}`$ represent different combinations of Coulomb interaction matrix elements, $`V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}`$, of the single-particle pseudospin states
$`V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_1^{},k_2^{},k_1,k_2}`$ $`=`$ $`{\displaystyle d^3\stackrel{}{r}_1d^3\stackrel{}{r}_2\psi _{\sigma _1^{},k_1^{}}^{}(\stackrel{}{r}_1)\psi _{\sigma _2^{},k_2^{}}^{}(\stackrel{}{r}_2)}`$ (10)
$`\times `$ $`{\displaystyle \frac{e^2}{ฯต|\stackrel{}{r}_1\stackrel{}{r}_2|}}\psi _{\sigma _1,k_1}(\stackrel{}{r}_1)\psi _{\sigma _2,k_2}(\stackrel{}{r}_2)`$ (11)
General expressions for the pseudospin dependent interactions $`W_{i,j}`$ are given in Table I in terms of the following matrix element combinations:
$`B_1^\pm `$ $`=`$ $`{\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,}),B_2^\pm ={\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,})`$ (12)
$`B_3^\pm `$ $`=`$ $`{\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,}),B_4^\pm ={\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,})`$ (13)
$`B_5^\pm `$ $`=`$ $`{\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,}),B_6^\pm ={\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,})`$ (14)
$`B_7^\pm `$ $`=`$ $`{\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,}),B_8^\pm ={\displaystyle \frac{1}{4}}(V_{,,,}\pm V_{,,,})`$ (15)
In (LABEL:b) we have omitted orbital guiding center indices for simplicity.
The many-electron state with pseudospin orientation $`\widehat{m}`$ is $`|\mathrm{\Psi }[\widehat{m}]=_{k=1}^{N_\varphi }c_{\widehat{m},k}^{}|0`$, where $`c_{\widehat{m},k}^{}`$ creates the single-particle state whose wavefunction is given in Eq. (6). We find that
$`e_{HF}(\widehat{m})`$ $``$ $`{\displaystyle \frac{\mathrm{\Psi }[\widehat{m}]|H|\mathrm{\Psi }[\widehat{m}]}{N_\varphi }}`$ (17)
$`=`$ $`{\displaystyle \underset{i=x,y,z}{}}\left(b_i{\displaystyle \frac{1}{2}}U_{\mathrm{๐},i}{\displaystyle \frac{1}{2}}U_{i,\mathrm{๐}}\right)m_i`$ (18)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i,j=x,y,z}{}}U_{i,j}m_im_j,`$ (19)
where
$$U_{i,j}=\frac{1}{N_\varphi }\underset{k_1,k_2=1}{\overset{N_\varphi }{}}\left(W_{i,j}^{k_1,k_2,k_1,k_2}W_{i,j}^{k_2,k_1,k_1,k_2}\right)$$
(20)
has direct and exchange contributions. Eq. (19) is the most general form for the HF energy of a pseudospin-1/2 QHF. The magnetic anisotropy of the particular QHF system is governed by the terms in (19) that are quadratic in the pseudospin magnetization $`m_i`$. The values of the coefficients $`U_{i,j}`$ depend on the nature of the crossing LLs and and can result in isotropic, easy-plane, or easy-axis quantum Hall ferromagnetism.
## IV Pseudospin matrix elements of the Coulomb interaction
This section contains the derivation of explicit expressions for the anisotropy energy coefficients $`U_{i,j}`$, assuming the zero-width quantum well model wavefunctions (1). Using the Fourier representation of the Coulomb interaction we write the pseudospin matrix elements as
$`V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_1^{},k_2^{},k_1,k_2}={\displaystyle \frac{1}{A}}{\displaystyle \underset{\stackrel{}{q}}{}}\delta _{q_y,k_1^{}k_1}\delta _{q_y,k_2^{}k_2}`$ (21)
$`\times `$ $`e^{iq_x(k_1^{}+k_1)/2}e^{iq_x(k_2^{}+k_2)/2}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(\stackrel{}{q})`$ (22)
and hence
$`{\displaystyle \frac{1}{N_\varphi }}{\displaystyle \underset{k_1,k_2=1}{\overset{N_\varphi }{}}}\left(V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_1,k_2,k_1,k_2}V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_2,k_1,k_1,k_2}\right)=`$ (23)
$`{\displaystyle \frac{N_\varphi }{A}}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(0){\displaystyle \frac{1}{A}}{\displaystyle \underset{\stackrel{}{q}}{}}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(\stackrel{}{q})`$ (24)
In the following we take $`\mathrm{}`$ as the unit of length and $`e^2/ฯต\mathrm{}`$ as the unit of energy. Then $`N_\varphi /A=1/2\pi `$ and Eq. (24) can be rewritten in the more transparent form
$`{\displaystyle \frac{1}{N_\varphi }}{\displaystyle \underset{k_1,k_2=1}{\overset{N_\varphi }{}}}\left(V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_1,k_2,k_1,k_2}V_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^{k_2,k_1,k_1,k_2}\right)=`$ (25)
$`{\displaystyle \frac{d^2\stackrel{}{q}}{(2\pi )^2}e^{q^2/2}\left[v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(0)v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(\stackrel{}{q})\right]}.`$ (26)
The first factor in square brackets in this equation originates from the Hartree contribution to the energy while the second factor originates from the exchange contribution.
In these equations $`v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}(\stackrel{}{q})`$ is a pseudospin-dependent effective 2D interaction which, because of the separability of the in-plane and out-of-plane degree-of-freedom terms in the single electron Schroedinger equation, is the product of two factors: the subband factor
$`v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^\mathrm{\Xi }(\stackrel{}{q})={\displaystyle _{\mathrm{}}^{\mathrm{}}}๐z_1{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐z_2e^{q|z_1z_2|}`$ (27)
$`\times `$ $`\lambda _{\xi (\sigma _1^{})}(z_1)\lambda _{\xi (\sigma _2^{})}(z_2)\lambda _{\xi (\sigma _1)}(z_1)\lambda _{\xi (\sigma _2)}(z_2)`$ (28)
and the in-plane term
$`v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^N(\stackrel{}{q})=e^{q^2/2}`$ (29)
$`\times `$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐x_1\varphi _{n(\sigma _1^{}),s(\sigma _1^{}),q_y/2}(x_1)\varphi _{n(\sigma _1),s(\sigma _1),q_y/2}(x_1)`$ (30)
$`\times `$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}๐x_2\varphi _{n(\sigma _2^{}),s(\sigma _2^{}),q_y/2}(x_2)\varphi _{n(\sigma _2),s(\sigma _2),q_y/2}(x_2)`$ (31)
For a general sample geometry the subband factor (28) has to be calculated numerically using the self-consistent LSDA wavefunctions. For our model bilayer system, however, we can obtain analytic expressions for $`v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^\mathrm{\Xi }(\stackrel{}{q})`$. In the case of $`\xi (\sigma )=\xi (\sigma )`$, Eqs. (1) and (28) give
$$v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^\mathrm{\Xi }=\frac{1}{2}\left[(1+r_\mathrm{\Delta }^2)+(1r_\mathrm{\Delta }^2)e^{dq}\right].$$
(32)
In the second case, i.e. $`\xi (\sigma )=\xi (\sigma )`$,
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^\mathrm{\Xi }={\displaystyle \frac{1}{2}}\left[(1+r_\mathrm{\Delta }^2)+(1r_\mathrm{\Delta }^2)e^{dq}\right],`$ (33)
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^\mathrm{\Xi }={\displaystyle \frac{1}{2}}\left[(1r_\mathrm{\Delta }^2)+(1+r_\mathrm{\Delta }^2)e^{dq}\right],`$ (34)
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^\mathrm{\Xi }={\displaystyle \frac{1}{2}}(1r_\mathrm{\Delta }^2)(1e^{dq}),`$ (35)
(36)
$`\mathrm{and}`$ (37)
(38)
$`v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^\mathrm{\Xi }=\eta {\displaystyle \frac{r_\mathrm{\Delta }}{2}}(1r_\mathrm{\Delta }^2)^{1/2}(1e^{dq})`$ (39)
$`\mathrm{if}\eta {\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{2}{}}}\left[\xi (\sigma _i^{})+\xi (\sigma _i)\right]=\pm 1.`$ (40)
For the in-plane factor in the effective interaction it is necessary to distinguish several cases. If the pseudospin-up and pseudospin-down levels have the same real-spin index and same orbit-radius quantum number, i.e. $`s(\sigma )=s(\sigma )`$ and $`n(\sigma )=n(\sigma )n`$, the effective interaction is independent of the pseudospin indices and we obtain from (3) and (31)
$$v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^N=\frac{2\pi }{q}\left[L_n(q^2/2)\right]^2,$$
(41)
where $`L_n(x)`$ is the Laguerre polynomial. For identical spins but different orbit-radius quantum numbers, i.e. $`s(\sigma )=s(\sigma )`$ and $`n(\sigma )n(\sigma )`$, we define $`n_<\mathrm{min}[n(\sigma ),n(\sigma )]`$ and $`n_>\mathrm{max}[n(\sigma ),n(\sigma )]`$. Then this factor in the effective interaction is
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}\left[L_{n(\sigma )}(q^2/2)\right]^2,`$ (42)
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}L_{n(\sigma )}(q^2/2)L_{n(\sigma )}(q^2/2),`$ (43)
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}{\displaystyle \frac{n_<!}{n_>!}}\left({\displaystyle \frac{q^2}{2}}\right)^{n_>n_<}\left[L_{n_<}^{n_>n_<}(q^2/2)\right]^2,`$ (44)
(45)
$`\mathrm{otherwise}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^N=0.`$ (46)
If the two pseudospin LLs have opposite real-spins then only scattering processes that conserve pseudospin (and therefore also real-spin) at each vertex will contribute to the anisotropy energy. For $`s(\sigma )=s(\sigma )`$ and $`n(\sigma )=n(\sigma )n`$ we obtain
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N=v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}\left[L_n(q^2/2)\right]^2,`$ (47)
(48)
$`\mathrm{otherwise}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^N=0.`$ (49)
Finally if the pseudospin LLs have opposite real-spins and different orbit-radius quantum numbers, the in-plane factor in the effective interactions is
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}\left[L_{n(\sigma )}(q^2/2)\right]^2,`$ (50)
$`v_{\sigma ,\sigma ,\sigma ,\sigma }^N={\displaystyle \frac{2\pi }{q}}L_{n(\sigma )}(q^2/2)L_{n(\sigma )}(q^2/2),`$ (51)
(52)
$`\mathrm{otherwise}v_{\sigma _1^{},\sigma _2^{},\sigma _1,\sigma _2}^N=0.`$ (53)
Eq. (28) for the subband factor and explicit expressions (41)-(53) for the in-plane factor in the effective interaction, together with Eqs. (26),(20),(LABEL:b) and Table I, provide a formal recipe to calculate the anisotropy coefficients $`U_{i,j}`$ for crossing LLs with any combination of quantum well subband, orbit radius and real-spin indeces. In the following section we use the explicit forms (40) and (32) for the subband factor to develope a pseudospin anisotropy classification scheme for our model bilayer system.
## V Magnetic anisotropy
The nature of the anisotropy energy is of qualitative importance for two-dimensional ferromagnets, including QHFs. Systems with easy-axis anisotropy, i.e., discrete directions at which the energy of the ordered state is minimized, have long range order at finite temperature and phase transitions in the Ising universality class. Systems with easy-plane anisotropy, i.e., a continuum of coplanar pseudospin magnetization orientations at which the energy of the ordered state is minimized, do not have long-range order but do have Kosterlitz-Thouless phase transitions at a finite temperature. In the isotropic case, all directions of pseudospin magnetization have identical energy, only the ground state has long-range order, and there are no finite-temperature phase transitions. Nevertheless, magnetization correlations become extremely long at low temperatures. The objective of this work is to predict which class of QHF occurs for a particular pair of crossing LLs.
We start our analysis by considering pseudospin LLs that belong to the same subband, i.e. $`\xi ()=\xi ()`$. In this case, only LLs with opposite real-spins can be aligned. Two examples of QHFs falling into this category, total filling factor $`\nu =1`$ with $`n()=n()=0`$, and $`\nu =2`$ with $`n()=1`$, $`n()=0`$, are illustrated in Fig. 2. The cases of $`n()n()`$ are realized when the ratio of the spin-splitting to LL separation is an integer. In GaAs, this ratio is only $`1/60`$ at perpendicular fields but can be tuned by tilting the magnetic field away from the normal to the 2D layer. For typical well widths orbital effects of the in-plane field, not included here, become important at the tilt angles where the coincidences of interest are realized and have to be accounted for to obtain correct values of the pseudospin anisotropy energy coefficients (20). However, recent work on AlAs quantum wells and InSb quantum wells with large Zeeman couplings have made the situation we study below, which assumes perpendicular magnetic field, accessible.
When opposite spin LLs cross, Eqs. (49) and (53) imply that all interactions $`B_i^\pm `$ with $`i>2`$ in (LABEL:b) vanish. Then the only non-zero anisotropy term is
$`U_{z,z}`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle _0^{\mathrm{}}}๐qe^{q^2/2}\left[L_{n()}(q^2/2)L_{n()}(q^2/2)\right]^2`$ (54)
$`\times `$ $`\left[(1+r_\mathrm{\Delta }^2)+(1r_\mathrm{\Delta }^2)e^{dq}\right].`$ (55)
Note that the Hartree energy contribution to anisotropy always vanishes when the crossing LLs share the same subband wavefunction. If the two pseudospin levels also have the same orbit-radius quantum number then Eq. (55) gives $`U_{z,z}=0`$ and the ferromagnetic state is isotropic. Physically, the result follows from the independence of the Coulomb interaction strength on real-spin. An important example of these isotropic QHFs occurs when $`n()=n()=0`$ and $`r_\mathrm{\Delta }=1`$, i.e., there is no tunneling between layers. This is the thoroughly studied single-layer $`\nu =1`$ QHF for which the HF theory ground state happens to be exact. We remark that quantitative estimates based on the HF mean-field theory presented here require corrections to quantum fluctuation effects in cases when the ordered pseudospin moment direction is not a good quantum number. A detail understanding of these corrections is one challenge for future experimental and theoretical work on QHFs.
For $`n()n()`$, Eq. (55) implies that $`U_{z,z}<0`$, making the $`z`$-axis the easy pseudospin orientation axis. Again, at $`r_\mathrm{\Delta }=1`$ our model reduces to that of a single layer 2D systems whose easy-axis anisotropy at even filling factors has been identified previously. In finite-thickness single quantum wells, a QHF with pseudospin LLs of the same subband but different real-spin and orbit-radius indices is also easy-axis. The magnitude of the anisotropy will decrease with layer thickness, as can be seen by comparing $`U_{z,z}`$ in (55) calculated for $`r_\mathrm{\Delta }=1`$ and $`r_\mathrm{\Delta }=0`$. (Note that the single-subband unbiased double well with finite tunneling, i.e. $`r_\mathrm{\Delta }=0`$, models a single layer system with an effective thickness $`d`$.)
We now turn to the crossing of LLs with different subband indices, for which the pseudospin anisotropy physics is richer. In Fig. 3 we show examples of $`n()=n()`$ bilayer QHFs for $`\nu =1`$ and $`\nu =2`$ based on same real-spin and opposite real-spin LLs respectively. Eqs. (20),(26),(40),(41), and Table I imply four non-zero anisotropy terms for $`n()=n()n`$ and $`s()=s()`$:
$`U_{z,z}=ur_\mathrm{\Delta }^2,`$ (56)
$`U_{x,x}=u(1r_\mathrm{\Delta }^2),`$ (57)
$`U_{x,z}=U_{z,x}=ur_\mathrm{\Delta }(1r_\mathrm{\Delta }^2)^{1/2},`$ (58)
(59)
$`\mathrm{where}`$ (60)
(61)
$`u={\displaystyle \frac{d}{2}}{\displaystyle \frac{1}{2}}{\displaystyle _0^{\mathrm{}}}๐qe^{q^2/2}\left[L_n(q^2/2)\right]^2\left(1e^{dq}\right).`$ (62)
First term in the expression for energy $`u`$ comes from the Hartree interaction, the second term represents exchange contribution which is always smaller than the Hartree energy in this case, i.e., $`u>0`$. For zero tunneling between layers ($`r_\mathrm{\Delta }=1`$), the only non-zero anisotropy energy component, $`U_{zz}=u`$, is positive leading to the easy-plane anisotropy of the QHF. In the absence of symmetry breaking fields (the linear pseudospin magnetization terms in (19)) the variational energy (19) is minimized when the pseudospins condense into a state magnetized at an arbitrary orientation within in the $`xy`$ plane. In this state, electronic charge is distributed equally between the layers (pseudospin angle $`\theta =0`$) minimizing the electrostatic energy; spontaneous interlayer phase coherence the physical counter part of pseudospin order in this case, lowers the total energy of the system by strengthening interlayer exchange interactions.
For $`r_\mathrm{\Delta }^2<1`$, Eqs. (62) imply the following quadratic terms in the HF total energy
$$\underset{i,j=x,y,z}{}U_{i,j}m_im_j=u\left[r_\mathrm{\Delta }m_z+(1r_\mathrm{\Delta }^2)^{1/2}m_x\right]^2,$$
(63)
i.e., the easy-plane is tilted from the $`xy`$ plane in the pseudospin space by angle $`\alpha =\mathrm{arctan}\left[(1r_\mathrm{\Delta }^2)^{1/2}/r_\mathrm{\Delta }\right]`$. The pseudospin basis states at different values of $`r_\mathrm{\Delta }`$ are related, however, by a unitary transformation, which corresponds precisely to a rotation about $`y`$-axis by angle $`\alpha `$, as seen from Eq. (1). The easy-plane where the above anisotropy energy is constant is, for any value of $`r_\mathrm{\Delta }`$, the plane of equal charge per layer.
When pseudospin-up and pseudospin-down states differ by more than their subband indices, by their spin indices for example, the two pseudospin basis sates at different $`r_\mathrm{\Delta }`$ are not related by a unitary transformation. In this case the magnetic anisotropy does depend on $`r_\mathrm{\Delta }`$. For example, consider the case $`\xi ()=\xi ()`$, $`n()=n()`$, and $`s()=s()`$. For opposite real-spin LLs, all anisotropy terms that include pseudospin non-conserving scattering processes drop out. The only non-zero energy term, $`U_{z,z}`$, has the same value as in (62). Hence, the QHF is isotropic in the unbiased ($`r_\mathrm{\Delta }=0`$) bilayer system while applying external bias ($`r_\mathrm{\Delta }>0`$) leads to easy-plane anisotropy in pseudospin space.
At this point let us make an experimentally important comment on the bilayer systems realized in wide single quantum wells. The difference between this sample geometry and the double quantum well with narrow (in our model infinitely narrow) layers is in the nature of the barrier responsible for the bilayer character of the electronic system. In wide quantum wells the barrier is soft, originating from Coulomb interactions among electrons in the well. Then the tunneling probability between layers is strongly dependent on the electron density and quantum well subband populations. This tunability makes wide single quantum wells an experimentally attractive alternative to double wells in studies of bilayer quantum Hall phenomena. The softness of the barrier can, however, lead to qualitatively important consequences for the ordered many-particle states. Translated into the pseudospin language, $`\mathrm{\Delta }_t`$ cannot be treated as an external one-body field acting on the pseudospin particles but, in general, will depend on the pseudospin orientation in the ordered ground state. For the pseudospin LLs discussed in the previous paragraph ($`\xi ()=\xi ()`$, $`n()=n()`$, and $`s()=s()`$) and for $`r_\mathrm{\Delta }=0`$, this effect can lead to an anisotropic QHF. LSDA calculations indicate that at low electron densities the anisotropy will be easy-plane while easy-axis anisotropy is more likely to develop at high densities. We make this remark to point out that not all results obtained for the double quantum well model are directly applicable to bilayers in wide single wells. In many cases the theoretical description of QHFs in wide single wells requires modifications of the idealized bilayer model to account for mixing of higher electrical subbands. The self-consistent LSDA for the growth direction single-particle orbitals is a particularly convenient, if somewhat ad hoc, method that allows any sample geometry to be studied while retaining the basic structure of the many-body HF formalism for QHFs.
In the remaining part of this section we consider pseudospin LLs with opposite subband indices and different orbit radius quantum numbers. At total filling factor $`\nu =3`$, for example, the pseudospin LLs will have the same real-spin while at $`\nu =4`$ opposite spin LLs can be aligned, as shown in Fig. 4. A common feature of the QHFs discussed below is the transition from a state with easy-axis anisotropy to a state with easy-plane anisotropy as $`r_\mathrm{\Delta }`$ and the layer separation $`d`$ are varied. For $`s()=s()`$, Eqs. (20),(26),(40),(46), and Table I give three non-zero anisotropy energies
$`U_{z,z}={\displaystyle \frac{r_\mathrm{\Delta }^2d}{2}}{\displaystyle \frac{1}{8}}{\displaystyle _0^{\mathrm{}}}๐qe^{q^2/2}`$ (64)
$`\times \{[L_{n()}(q^2/2)L_{n()}(q^2/2)]^2(1+e^{dq})`$ (65)
$`+r_\mathrm{\Delta }^2[L_{n()}(q^2/2)+L_{n()}(q^2/2)]^2(1e^{dq})\},`$ (66)
$`U_{x,x}=U_{y,y}={\displaystyle \frac{1}{4}}{\displaystyle _0^{\mathrm{}}}๐qe^{q^2/2}{\displaystyle \frac{n_<!}{n_>!}}\left({\displaystyle \frac{q^2}{2}}\right)^{n_>n_<}`$ (67)
$`\times \left[L_{n_<}^{n_>n_<}(q^2/2)\right]^2\left(1r_\mathrm{\Delta }^2\right)\left(1e^{dq}\right),`$ (68)
(69)
$`\mathrm{where}`$ (70)
(71)
$`n_<\mathrm{min}[n(),n()],n_>\mathrm{max}[n(),n()].`$ (72)
The HF total energy contributions that are quadratic in the pseudospin magnetization components can be grouped as
$$\underset{i,j=x,y,z}{}U_{i,j}m_im_j=(U_{z,z}U_{x,x})m_z^2+U_{x,x}.$$
(73)
(Recall that $`\widehat{m}`$ is a unit vector, i.e. $`m_x^2+m_y^2=1m_z^2`$.) From Eq. (73) we obtain that for $`U_{z,z}U_{x,x}<0`$ the QHF has easy-axis anisotropy while for $`U_{z,z}U_{x,x}>0`$ the system is an easy-plane ferromagnet. At the critical layer separation $`d=d^{}`$, obtained from the condition $`U_{z,z}=U_{x,x}`$, the magnetic anisotropy vanishes and a fine-tuned isotropy is achieved. It follows from Eqs. (72) that $`d^{}`$ is finite for all values of $`r_\mathrm{\Delta }`$.
For pseudospin LLs with $`s()=s()`$, the anisotropy energy components $`U_{x,x}`$ and $`U_{y,y}`$ vanish and the critical layer separation $`d^{}`$ corresponds to $`U_{z,z}=0`$, where $`U_{z,z}`$ is given by the same expression as in the $`s()=s()`$ case (see Eq. (72)). Since $`U_{z,z}<0`$ at $`r_\mathrm{\Delta }=0`$, the critical separation $`d^{}`$ diverges in the absence of external bias, i.e., easy-axis pseudospin anisotropy does not exist for any layer separation. For $`r_\mathrm{\Delta }>0`$ the transition between easy-plane and easy-axis anisotropy occurs at finite $`d`$ as for the $`s()=s()`$ pseudospin LLs.
In Figs. 5(a) and 5(b) we show the magnetic anisotropy phase diagrams in the $`d`$-$`r_\mathrm{\Delta }`$ plane calculated for $`\nu =3`$ and $`\nu =4`$ QHFs (see Fig. 4). Since the layer separation is in units of magnetic length, these figures imply that transition between easy-axis and easy-plane anisotropies at a given filling factor can be induced in one physical sample by changing the density of the 2D electron system. High electron densities would correspond to the easy-plane region, and low densities to the easy-axis region. Note that these numerical results confirm the general remark made above, since the critical layer diverges as $`r_\mathrm{\Delta }0`$ for $`\nu =4`$ while it remains finite for $`\nu =3`$.
## VI Symmetry breaking fields
The pseudospin orientation in a QHF ground state is determined by minimizing the variational total energy (19). In the absence of energy terms that are linear in pseudospin magnetization components, the HF ordered states spontaneously break continuous SU(2) or U(1) symmetry in the case of isotropic or easy-plane QHFs respectively, and the discrete symmetry between pseudospin-up and pseudospin-down orientations in the case of easy-axis QHFs. In this section we take into account external and internal potentials which contribute to the linear terms in the HF total energy and comment on the pseudospin reversal that can be triggered by adjusting these symmetry breaking fields. We focus on an case which we feel is particularly appropriate for experimental study by considering the ordered $`\nu =3`$ quantum Hall state. Similar considerations would apply for all classes of QHFs discussed in this paper.
The pseudospin LLs in the bilayer $`\nu =3`$ QHF (see Fig. 4) have opposite subband indices and orbit radius quantum numbers $`n=0`$ and $`n=1`$, respectively. We call the \[$`\xi =1`$,$`n=0`$,$`s=+1/2`$\] LL the pseudospin-up state and the \[$`\xi =1`$,$`n=1`$,$`s=+1/2`$\] LL the pseudospin-down state. With this definition, the one-body potentials in (19) can be written as
$`b_z`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[(\mathrm{\Delta }_V^2+\mathrm{\Delta }_t^2)^{1/2}\mathrm{}\omega _c+I_FI_{H,z}\right],`$ (74)
$`b_x`$ $`=`$ $`{\displaystyle \frac{1}{2}}I_{H,x},`$ (75)
$`b_y`$ $`=`$ $`0.`$ (76)
The effective field $`I_F`$ is the difference between pseudospin-up and pseudospin-down particle exchange energy with electrons in the fully occupied \[$`\xi =1`$,$`n=0`$,$`s=+1/2`$\] LL, i.e,
$`I_F`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^{\mathrm{}}}dqe^{q^2/2}\{{\displaystyle \frac{q^2}{2}}[1+r_\mathrm{\Delta }^2+(1r_\mathrm{\Delta }^2)e^{dq}]`$ (77)
$``$ $`(1r_\mathrm{\Delta }^2)(1e^{dq})\}.`$ (78)
For $`r_\mathrm{\Delta }>0`$, electrons in the \[$`\xi =1`$,$`n=0`$,$`s=\pm 1/2`$\] LLs produce also an electrostatic field, represented by $`I_{H,z}`$ and $`I_{H,x}`$ in (76), which screens the external bias potential. Since this effective field favors occupation of a particular layer rather than a particular pseudospin state it couples to $`z`$ and $`x`$ components of the pseudospin operator. The energy inbalance between the two layers produced by $`\stackrel{}{I}_H`$ is $`2dr_\mathrm{\Delta }`$ which, together with Eq. (1), gives
$`I_{H,z}`$ $`=`$ $`2dr_\mathrm{\Delta }^2,`$ (79)
$`I_{H,x}`$ $`=`$ $`2dr_\mathrm{\Delta }(1r_\mathrm{\Delta }^2)^{1/2}.`$ (80)
In the expression for the HF total energy (19), we included explicitly the contribution to the symmetry breaking fields which results from Coulomb interactions between electrons in the pseudospin LLs. For the $`\nu =3`$ QHF we are considering here, only $`U_{\mathrm{๐},z}`$ and $`U_{z,\mathrm{๐}}`$ energies are non-zero:
$`U_{\mathrm{๐},z}=U_{z,\mathrm{๐}}`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle _0^{\mathrm{}}}๐qe^{q^2/2}\left[1+r_\mathrm{\Delta }^2+(1r_\mathrm{\Delta }^2)e^{dq}\right]`$ (81)
$`\times `$ $`(q^2q^4/4).`$ (82)
In the bilayer system with no tunneling ($`r_\mathrm{\Delta }=1`$), $`I_{H,x}=0`$ and the total symmetry breaking field, $`b^{}b_zU_{\mathrm{๐},z}`$, is oriented along the $`z`$ pseudospin direction and is given by
$`b^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[(\mathrm{\Delta }_V^2+\mathrm{\Delta }_t^2)^{1/2}\mathrm{}\omega _c+\sqrt{\pi /2}/22d`$ (83)
$``$ $`\sqrt{\pi /2}/8].`$ (84)
In Figs. 6(a)-(c) we plot the pseudospin evolution with effective field $`b^{}`$ for the three anisotropy regimes of a $`\nu =3`$ QHF with $`r_\mathrm{\Delta }=1`$. At the phase boundary between easy-axis and easy-plane anisotropies, the $`\nu =3`$ QHF is isotropic and the pseudospin reverses abruptly at $`b^{}=0`$ (see Fig. 6(a)). In the easy-plane anisotropy regime, ($`U_{z,z}U_{x,x}>0`$), the pseudospin evolves continuously with $`b^{}`$ as illustrated in Fig. 6(b) reaching alignment with $`b^{}`$ at $`|b^{}|U_{z,z}U_{x,x}`$. For the easy-axis anisotropy case ($`U_{z,z}U_{x,x}<0`$), the HF energy has two local minima at $`m_z=\pm 1`$ when $`|b^{}|<|U_{z,z}U_{x,x}|`$. The pseudospin-up and pseudospin-down polarized states are separated by an energy barrier which results in the hysteretic pseudospin-reversal behavior shown in Fig. 6(c).
In bilayer systems with non-zero tunneling, pseudospin reversal follows a more complicated pattern in which the competition between $`x`$ and $`z`$ components of the symmetry breaking field plays an important role. In general, the pseudospin will rotate in the $`x`$-$`z`$ plane, i.e. $`m_x=\sqrt{1m_z^2}`$. Since the derivative of the HF energy with respect to $`m_z`$ diverges at $`m_z=\pm 1`$ due to the $`I_{H,x}`$ term the pseudospin will never align completely with the $`z`$-axis when $`r_\mathrm{\Delta }<1`$.
## VII Summary
In the strong magnetic field limit, the physics of high mobility two-dimensional electron systems is usually dominated by electron-electron interactions except at integer filling factors, where the single-particle physics responsible for the gap between Landau levels assumes the dominant role. When external parameters are adjusted so that two or more Landau levels simultaneously approach the chemical potential, the integer filling factor case is less exceptional, interaction effects are always strong, and uniform density broken symmetry ground states analogous to those in conventional ferromagnets are common. In this paper we have discussed how the nature of these states depends on the character of the nearly degenerate Landau levels. Our attention is restricted to the case where only two Landau levels are close to the chemical potential and we distinquish these crossing Landau levels by introducing a pseudospin degree of freedom.
Using Hartree-Fock variational wavefunctions, we are able to derive an explicit expression for the dependence of ground state energy on pseudospin orientation for crossing LLs with any combination of quantum well subband, orbit radius and real-spin degree-of-freedom quantum numbers. As in conventional magnetic systems, qualitative differences exist between the physical properties of isotropic (Heisenberg) systems with no dependence of energy on pseudospin orientation, easy-axis (Ising) systems with discrete prefereed pseudospin orientations, and (XY) easy-plane systems for which the minimum is achieved simultaneously for a plane of orientations. Our mean-field results predict which class of pseudospin quantum Hall ferromagnet occurs in different circumstances. We focus on a model commonly used for bilayer quantum Hall systems in which the finite width of both quantum wells is neglected. The external parameters of the model are the Zeeman coupling strength, the bias potential between the wells $`\mathrm{\Delta }_V`$, and the single particle splitting due to interlayer tunneling $`\mathrm{\Delta }_t`$. In the limit $`\mathrm{\Delta }_t=0`$, this model applies to a single quantum well when the crossing Landau levels have the same subband wavefunction, i.e., are in the same quantum well. Classification predictions for this model as a function of layer separation $`d`$ and the ratio $`r_\mathrm{\Delta }=\mathrm{\Delta }_V/(\mathrm{\Delta }_V^2+\mathrm{\Delta }_t^2)^{1/2}`$ are summarized in the diagram 7.
For the single quantum well case we find isotropic behavior when the orbit radius quantum numbers of the crossing Landau levels are identical, and easy-axis behavior otherwise. In general when the crossing Landau levels have identical subband wavefunctions, the nature of the pseudospin anisotropy does not depend on the paramters $`d`$ and $`r_\mathrm{\Delta }`$. For different subband wavefunctions the pseudospin anisotropy can vary in the $`dr_\mathrm{\Delta }`$ plane. Particularly intriqueing is the case of crossing LLs with different subband and orbit radius quantum numbers where, at a given filling factor, the system can undergo a quantum phase transition from an easy-axis to easy-plane QHF. At the phase boundary the pseudospin anisotropy vanishes and a fine-tuned isotropy is achieved. The critical values of $`d`$ and $`r_\mathrm{\Delta }`$ are experimentally accesible and may be accompanied by observable changes in pseudospin reversal properties as external parameters are varied.
## Acknowledgments
This work was supported by the National Science Foundation under grants DMR-9623511, DMR-9714055, and DGE-9902579 and by the Grant Agency of the Czech Republic under grant 202/98/0085. |
warning/0003/hep-ph0003109.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The Standard Model (SM) of strong (QCD) and electroweak (EW) interactions has been the most successful framework to describe the phenomenology of high energy physics. However, one of the fundamental building blocks (the Higgs boson) still lacks experimental confirmation. On the other hand, the SM still suffers from some theoretical deficiencies, most notably the hierarchy problem. Several extensions of the SM have been proposed to solve these problems; in this note we will concentrate on its supersymmetric (SUSY) extension, more specifically to the $`R`$-parity conserving Minimal Supersymmetric Standard Model (MSSM) .
The top quark, due to its large mass, could play a central role in the search of physics beyond the SM. On one hand it could decay to non-standard particles, on the other hand, due to its large Yukawa coupling, the effects of the Spontaneous Symmetry Breaking Sector are expected to be larger than for any other particle of the model. In the MSSM, these effects are reinforced by the presence of the SUSY partners of the top quark and Higgs bosons. Moreover, the new parameters appearing in the MSSM can have complex phases, and new sources of CP-violation phenomena can appear.
One should also bear in mind that, before the commissioning of TESLA, the LHC will be producing data which might turn out to include some physics beyond the SM, and we must be prepared for whatever the high energy physics scenario would be for the running of TESLA.
Here we present a review of the effects of SUSY in the top-quark phenomenology. First we review in section 2 the top-quark decay within the framework of the MSSM; then, in section 3, we present the MSSM effects on the top quark pair production in $`e^+e^{}`$ collisions, both in the presence and absence of CP-violating couplings.
## 2 Top-quark decays
The existence of SUSY could affect the total top-quark decay width in two ways. First of all through unexpected radiative corrections to the standard top quark decay process $`tW^+b`$. Second, some of the SUSY particles could be lighter than the top quark itself, thus providing new channels in which the top quark could decay.
Concerning the standard top-quark decay, we point out that its branching ratio $`\text{BR}(tW^+b)`$ is not so severely constrained by the observed top quark production cross section as one could naively think at first sight . In the MSSM the total observed cross section can be written, schematically, as
$$\begin{array}{ccc}\hfill \sigma _{\mathrm{obs}}& =& ๐q๐\overline{q}\sigma (q\overline{q}t\overline{t})\times |\text{BR}(tW^+b)|^2\hfill \\ & +& ๐q๐\overline{q}\sigma (q\overline{q}\stackrel{~}{g}\overline{\stackrel{~}{g}})\times |\text{BR}(\stackrel{~}{g}t\overline{\stackrel{~}{t}}_1)|^2\times |\text{BR}(tW^+b)|^2\hfill \\ & +& ๐q๐\overline{q}\sigma (q\overline{q}\stackrel{~}{b}_a\overline{\stackrel{~}{b}}_a)\times |\text{BR}(\stackrel{~}{b}_at\chi _1^{})|^2\times |\text{BR}(tW^+b)|^2+\mathrm{}\hfill \end{array}$$
Here $`๐q`$ represents the integration over the quarks Parton Distribution Functions, and sumation over quark flavours. In the SM only the first line of this equation is present. It follows that, in the MSSM, our present ignorance of the SUSY parameters prevents us from performing a detailed calculation of the $`t\overline{t}`$ production cross section as well as from putting a strict limit on $`\text{BR}(t`$โnewโ$`)`$ โ the branching ratio of the top quark into new physics. It should also be clear that the observed cross section in Eqn. (2) refers not only to the standard $`bWbW`$ events, but to all kind of final states that can mimic them. Thus, effectively, we should substitute $`\text{BR}(tXb)`$ in that formula for $`\text{BR}(tW^+b)`$, and then sum the cross section over $`X`$, where $`X`$ is any state that leads to an observed pattern of leptons and jets similar to those resulting from $`W`$-decay. In particular, $`X=H^\pm `$ would contribute (see below) to the $`\tau `$-lepton signature, if $`\mathrm{tan}\beta `$ is large enough. Similarly, there can be direct top quark decays into SUSY particles that could mimic the SM decay of the top quark . The only restriction is an approximate lower bound $`\text{BR}(tW^+b)\text{ }\stackrel{>}{}\text{ }4050\%`$ in order to guarantee the purported standard top quark events at the Tevatron . Thus, from these considerations it is not excluded that the non-SM branching ratio of the top quark, $`\text{BR}(t`$โnewโ$`)`$, could be comparable to the SM one โ or at least not necessarily much smaller.
The $`e^+e^{}`$ Linear Collider (LC) provides an excellent tool to test the various top quark partial decays widths. The total top quark decay width can be measured by means of a threshold scan, in a model independent way . Also the clean environment allows for a high prospect for detecting exclusive rare decay channels.
### 2.1 The top-quark standard decay width
In the SM the top-quark decays into a $`W^+`$ gauge boson and a bottom quark. The tree-level prediction for this partial decay width is (for $`m_t=175GeV`$)
$`\mathrm{\Gamma }_{\mathrm{SM}}^0(tW^+b)`$ $`=`$ $`\left({\displaystyle \frac{G_F}{8\pi \sqrt{2}}}\right){\displaystyle \frac{|V_{tb}|^2}{m_t}}\lambda ^{1/2}(1,m_b^2/m_t^2,M_W^2/m_t^2)`$ (1)
$`\times [M_W^2(m_t^2+m_b^2)+(m_t^2m_b^2)^22M_W^4]1.55\text{ GeV}`$
where $`\lambda ^{1/2}(1,x^2,y^2)`$ is the usual Kรคllen function. As the width turns out to be much larger than the typical scale of non-perturbative QCD effects $`\mathrm{\Lambda }_{QCD}`$, it can be conceived as an effective infrared cut-off. This means that the top quark, due to its large mass, has time to weakly decay before strong hadronization processes come into play. And for this reason perturbative computations in top quark physics are reliable.
In this spirit the SM quantum corrections to the standard top quark decay width have been performed. The short-distance QCD effects have been computed up to two-loop level and they amount to a correction of $`10\%`$ with respect to the tree-level width. The electroweak (EW) SM radiative correction is also available , but this contribution is below $`+2\%`$ in a scheme where the tree-level width is parametrized in terms of the Fermi constant $`G_F`$ โ as in Eqn.(1). In this scheme the electroweak corrections are minimized both in the SM and in the MSSM because the set of universal contributions โ viz. those encoded in the parameter $`\mathrm{\Delta }r`$ โ cancel out.
The MSSM may furnish extra (perturbative) quantum effects on the standard decay width of the top quark through the one-loop corrections mediated by non-SM particles. They have been computed in <sup>2</sup><sup>2</sup>2The corresponding corrections in the general 2HDM can be found in Refs. . The particularization of these Higgs effects in the MSSM is studied in . and can be of two types, electroweak and strong. The SUSY-EW quantum corrections are negative (as the standard QCD ones) and vary from $`1\%`$ to $`10\%`$, depending on the choice of the various SUSY parameters, and, especially of $`\mathrm{tan}\beta `$. The corrections due to additional Higgs particle exchange (i.e. the MSSM Higgs effects after subtracting the corresponding SM limit of the MSSM Higgs sector) are at most of $`0.1\%`$ due to the severe constraints that SUSY imposes on the MSSM Higgs sector.
The SUSY-QCD corrections, mediated by gluinos and squarks, have also been found to be negative in most of the parameter space , though they are in general smaller than the SUSY-EW ones โ namely, around a few % level โ and they are independent of $`\mathrm{tan}\beta `$.
In Fig. 1 we show the total SUSY (electroweak and QCD) corrections to this decay width for typical values of the SUSY spectrum. The upshot is that the total SUSY corrections to $`\mathrm{\Gamma }(tW^+b)`$ go in the same direction as the standard QCD ones. For large values of $`\mathrm{tan}\beta `$ they can typically yield an effect about half the size of the QCD corrections, thus providing an additional (potentially measurable) decrease of the tree-level value of the standard width (1).
### 2.2 Top-quark decay into charged Higgs
If the charged Higgs is light enough the top quark will also decay through the process $`tH^+b`$. This decay has been subject of interest since very early in the literature . If $`\mathrm{tan}\beta `$ is large or small enough the tree-level prediction for the partial decay width $`\mathrm{\Gamma }^0(tH^+b)`$ is comparable to the standard one (1). In fact $`\mathrm{\Gamma }^0(tH^+b)`$ presents a minimum at the point $`\mathrm{tan}\beta =\sqrt{m_t/m_b}6`$ and grows for larger or smaller values of $`\mathrm{tan}\beta `$ (see Fig. 2). Remarkably enough, this process turns out to be extremely sensitive to radiative corrections of all kinds. On one hand, the standard QCD corrections are quite large. They are negative and for $`\mathrm{tan}\beta \text{ }\stackrel{>}{}\text{ }10`$ they saturate around the value $`58\%`$ . On the other hand, the full set of the MSSM radiative corrections, at the one-loop level, can also be very important. They have been computed in , and more recently the general 2HDM corrections became also available .
In the case of the EW corrections one needs to define renormalization prescriptions also for the non-SM parameters that appear in these decays, in particular for the highly relevant parameter $`\mathrm{tan}\beta `$. The renormalization counterterm for $`\mathrm{tan}\beta `$ can be fixed in many different ways. The actual corrections will depend on the particular definition, but not so the value of the physical observable, of course. In our case we found it practical to define $`\mathrm{tan}\beta `$ through the condition that the partial decay width $`\mathrm{\Gamma }(H^+\tau ^+\nu _\tau )`$ does not receive radiative corrections. This is a good choice for the scenario under study, since this is the dominant decay of a light charged Higgs boson (i.e. $`M_{H^\pm }<m_t`$) provided $`\mathrm{tan}\beta `$ is of order $`1`$ or above: $`\mathrm{tan}\beta >\sqrt{m_c/m_s}\text{ }\stackrel{>}{}\text{ }2`$. Under this renormalization prescription, and for moderate or large $`\mathrm{tan}\beta \text{ }\stackrel{>}{}\text{ }10`$, the bulk of the SUSY quantum corrections are known to stem from the finite threshold corrections to the bottom mass counterterm. The relevant effects are triggered by $`R`$-odd particles entering the bottom self-energy, and can be cast as follows
$`\left({\displaystyle \frac{\delta m_b}{m_b}}\right)_{\mathrm{S}\mathrm{QCD}}`$ $`=`$ $`{\displaystyle \frac{2\alpha _s(m_t)}{3\pi }}m_{\stackrel{~}{g}}M_{LR}^bI(m_{\stackrel{~}{b}_1},m_{\stackrel{~}{b}_2},m_{\stackrel{~}{g}})`$
$``$ $`{\displaystyle \frac{2\alpha _s(m_t)}{3\pi }}m_{\stackrel{~}{g}}\mu \mathrm{tan}\beta I(m_{\stackrel{~}{b}_1},m_{\stackrel{~}{b}_2},m_{\stackrel{~}{g}}),`$
$`\left({\displaystyle \frac{\delta m_b}{m_b}}\right)_{\mathrm{S}\mathrm{EW}}`$ $`=`$ $`{\displaystyle \frac{h_th_b}{16\pi ^2}}{\displaystyle \frac{\mu }{m_b}}m_tM_{LR}^tI(m_{\stackrel{~}{t}_1},m_{\stackrel{~}{t}_2},\mu )`$
$``$ $`{\displaystyle \frac{h_t^2}{16\pi ^2}}\mu A_t\mathrm{tan}\beta I(m_{\stackrel{~}{t}_1},m_{\stackrel{~}{t}_2},\mu ),`$
$`I(m_1,m_2,m_3)`$ $``$ $`16\pi ^2iC_0(0,0,m_1,m_2,m_3)`$ (2)
$`=`$ $`{\displaystyle \frac{m_1^2m_2^2\mathrm{ln}{\displaystyle \frac{m_1^2}{m_2^2}}+m_2^2m_3^2\mathrm{ln}{\displaystyle \frac{m_2^2}{m_3^2}}+m_1^2m_3^2\mathrm{ln}{\displaystyle \frac{m_3^2}{m_1^2}}}{(m_1^2m_2^2)(m_2^2m_3^2)(m_1^2m_3^2)}},`$
where $`C_0`$ is the three-point โt Hooft-Passarino-Veltman function , and the rightmost expressions hold for sufficiently large $`\mathrm{tan}\beta `$. Several important consequences can be derived already from the approximate expressions (2). The first one corresponds to the sign of the corrections. The sign of the SUSY-QCD corrections is opposite to that of the higgsino mass parameter $`\mu `$, whereas the sign of the SUSY-EW corrections is given by the product of $`\mu `$ and the soft-SUSY-breaking trilinear coupling $`A_t`$. Second, both kind of corrections grow linearly with $`\mathrm{tan}\beta `$. A third, and important, observation is that, if we scale all the dimensionful parameters of Eqn. (2) by a factor $`\lambda `$, the $`\lambda `$-dependence drops out in the final expression. This means that raising the scale of SUSY breaking does not reduce the effects of the radiative corrections. Notice, however, that this consideration amounts to scale up also the trilinear coupling $`A_t`$ as well as the higgsino parameter $`\mu `$. Therefore, it may lead to unwanted fine-tuning effects in at least two important sectors of the MSSM: in the Higgs and squark sectors. Notwithstanding, if one does not stretch out the ranges of the parameters up to unreasonable limits (i.e. much beyond $`1`$ TeV or so), an important consequence can be derived without disrupting the natural structure of the model, to wit: that the $`R`$-odd particles whose masses are above the EW scale can effectively display, in the presence of Yukawa couplings, a non-decoupling behaviour. And this behaviour is triggered by the existence of explicit soft SUSY-breaking terms in combination with the spontaneous breaking of the gauge symmetry. Obviously, this is a very important feature as it could produce visible radiative corrections for this decay.<sup>3</sup><sup>3</sup>3We should like to say that this feature is not limited to just the high energy process under consideration, but it applies equally well to some low-energy processes, e.g. in B-meson decays . Recently these finite threshold effects have been further refined in the literature and they have been re-summed to all orders .
In the case of the gluino mass dependence there is another trait that we wish to remark. Even without scaling up the rest of the SUSY parameters, the SUSY-QCD corrections to $`\mathrm{\Gamma }^0(tH^+b)`$ exhibit a local, and lengthy sustained, maximum around $`m_{\stackrel{~}{g}}\text{ }\stackrel{>}{}\text{ }300\text{ GeV}`$. Only for gluino masses well above the TeV scale (for fixed values of the squark masses) do these corrections eventually decouple . As for the corrections due to Higgs bosons loops, of which there are quite a few, we find that they are entirely negligible compared to the yield from $`R`$-odd particles.<sup>4</sup><sup>4</sup>4This effect is due to the restrictions that SUSY imposes to the form of the Higgs bosons potential. In the unrestricted 2HDM the Higgs bosons loop corrections can also be important .
Figure 2a presents a summary of the main results. Here we have plotted the partial decay width $`\mathrm{\Gamma }(tH^+b)`$ as a function of $`\mathrm{tan}\beta `$, for the two different scenarios that have been identified. In the first one ($`\mu <0`$) the SUSY-QCD corrections are opposite in sign to the QCD ones, thus canceling partially (or even totally) the SM strong corrections. In the second one ($`\mu >0`$) the SUSY-QCD corrections have the same sign as the standard QCD ones, so reinforcing the large negative corrections. In both cases we have fixed $`\mu A_t<0`$, which is the overall sign which makes allowance for the low energy data on radiative $`B`$ meson decays to be compatible with the existence of a light charged Higgs below the top quark mass . This fixes the SUSY-EW corrections (2) to be positive. In Fig. 2b we present the relative corrections induced by each sector of the MSSM, in the $`\mu <0`$ scenario.
The results shown in Fig. 2 can hardly be overemphasized. Whereas the QCD prediction for the partial decay width states that this is always significantly smaller than the standard partial decay width, in the $`\mu <0`$ scenario the charged Higgs partial decay width is equal to the standard one for $`\mathrm{tan}\beta 50`$, and it is rapidly increasing. In this scenario the presence of charged Higgs in top quark decays is significantly greater than the QCD prediction, and thus the experimental discovery reach of charged Higgs in top quark decays can be larger than expected. On the other hand in the $`\mu >0`$ scenario the discovery reach is decreased with respect to the standard one. Thus the excluded region in the $`\mathrm{tan}\beta M_{H^\pm }`$ plane due to the (up to now) unsuccessful search of charged Higgs bosons in top quark decays depends drastically on the value of the rest of parameters in the MSSM .
In Fig. 2b we see clearly the close-to-linear behaviour of the leading SUSY contributions (2), and we also see that the Higgs-boson mediated contributions ($`\delta _{EW}`$ in the plot) really play a marginal role. Interestingly enough we see that for $`\mathrm{tan}\beta 35`$ the SUSY-QCD corrections cancel the standard QCD ones, and thus, although the larger corrections are due to the strong interaction sector, the only radiative corrections that are left are the SUSY-EW ones.
The SUSY radiative corrections above $`\mathrm{tan}\beta 35`$ (at one-loop) can easily reach values of
$`\delta _{\mathrm{S}\mathrm{EW}}+30\%,\delta _{\mathrm{S}\mathrm{QCD}}+80\%(\mu <0,A_t>0,M_{SUSY}100200\text{ GeV}),`$
$`\delta _{\mathrm{S}\mathrm{EW}}+20\%,\delta _{\mathrm{S}\mathrm{QCD}}40\%(\mu >0,A_t<0,M_{SUSY}500\text{ GeV}).`$ (3)
Negative corrections for the SUSY-EW corrections of the same absolute values are possible provided $`\mu A_t>0`$. We have singled out different sparticle spectra for the two scenarios in order to avoid total corrections greater than $`100\%`$ when they are added to the standard QCD corrections.
### 2.3 FCNC top-quark decays
Flavour Changing Neutral Current (FCNC) decays of the top quark are one-loop induced processes. They are such rare events in the SM , with branching ratios at the level of $`10^{10}10^{15}`$ depending on the particular channel, that its presence at detectable levels would clearly indicate the presence of new physics. The question is whether the presence of SUSY particles could enhance these partial decay widths up to the visible level. The partial FCNC decay width into a weak vector boson $`\mathrm{\Gamma }(tcV)`$ ($`V=Z,\gamma `$) undergoes some enhancement , however it is still of the order of $`10^{12}10^{13}`$ in most of the parameter space, thus being far away of the detection level. The gluon channel ($`tcg`$) is the most gifted one in the SM, but it is nevertheless too small to be detectable ($`10^{10}`$). This mode, however, has recently been analyzed in great detail in the MSSM and one finds that its branching ratio can be close to the visible threshold for the future high luminosity machines such as the LHC and the LC (see below). Finally, the top quark decaying into neutral Higgs particles ($`tch`$, $`h=h^0`$, $`H^0`$, $`A^0`$) has also been shown to benefit from large enhancements in the MSSM framework . In this respect we recall that in the MSSM the channel in which the lightest Higgs boson is involved ($`tch^0`$) is always kinematically open because $`m_{h^0}130GeV<m_t`$. Hereafter we will concentrate on the two decays, $`tcg`$ and $`tch^0`$, because the overall analysis shows that they are the most efficient FCNC decays in their respective modalities. Of course in SUSY models beyond the MSSM, such as models without $`R`$-parity, there could be other kind of competing FCNC top quark decays,<sup>5</sup><sup>5</sup>5See e.g. Refs. for some recent works on the subject. but here we shall stick all the time to the MSSM.
FCNC processes can be induced through SUSY-EW charged current interactions. These proceed through the same mixing matrix elements as in the SM: the Cabibbo-Kobayashi-Maskawa mixing matrix. But in addition it could happen that the squark mass matrix squared is not proportional to the quark-mass-matrix squared. In this case the squark mass eigenstates would not coincide with the quark mass states, and as a consequence tree-level FCNC would appear in the quark-squark-gaugino/higgsino interactions. This mixing appears as non-flavour diagonal mass matrix elements in the squark mass matrix squared
$$(M_{LL}^2)_{ij}=m_{ij}^2\delta _{ij}m_im_j,(ij),$$
(4)
where $`i,j`$ represent squarks of any generation, and $`m_{\{i,j\}}`$ is the mass corresponding to the diagonal entries in the matrix. In the MSSM these kind of mixing terms in the Left-chiral sector are naturally generated through the Renormalization Group evolution of the soft-SUSY-breaking squark masses down to the EW scale . This is the reason why we just singled out the $`LL`$ mixing component in Eqn. (4). Flavour mass mixing terms for the corresponding Right-chiral squarks are allowed, but they do not appear naturally in the GUT frameworks. Moreover its presence is not essential because it does not change the order of magnitude of the results obtained with only flavour-mixing between Left-chiral squarks . The mixing terms $`\delta _{ij}`$ are restricted by the low energy data on FCNC processes . These limits were computed using the mass-insertion approximation, so they should be taken as order of magnitude limits. Recently it has been shown that the full computation of some FCNC process can give results which differ substantially from the mass-insertion approximation ones .
To assess the size of the FCNC top quark decay rates we use the fiducial ratio
$$B(tch)\frac{\mathrm{\Gamma }(tcX)}{\mathrm{\Gamma }(tbW^+)},$$
(5)
for both $`X=g,h^0`$. The typical values of this ratio lie in the ballpark of $`10^8`$ for the SUSY-EW contributions in the regime of large $`\mathrm{tan}\beta `$ ($`30\text{ }\stackrel{<}{}\text{ }\mathrm{tan}\beta \text{ }\stackrel{<}{}\text{ }50)`$ and for a SUSY spectrum around $`200\text{ GeV}`$. This is already five orders of magnitude larger than the corresponding processes in the SM. The SUSY-QCD contributions, which appear when the $`\delta _{23}`$ in Eqn. (4) is non-zero for up-type squarks, are typically around two orders of magnitude larger, and they exhibit a slow decoupling as a function of the gluino mass, so even for gluinos as heavy as $`m_{\stackrel{~}{g}}500\text{ GeV}`$ the ratio $`B(tch^0)`$ can reach the level of $`10^5`$. This large value is due to the strong nature of the gluino-mediated interactions, but not less to the fact that present bounds on $`\delta _{23}`$ are rather poor. Also the various decays are sensitive to both: the higgsino mass parameter $`\mu `$ and the soft-SUSY-breaking trilinear coupling $`A_t`$.
In Figs. 3a and b we present the result of maximizing the ratio (5) for the SUSY-EW and SUSY-QCD contributions respectively . These plots have been obtained by performing a full scan of the MSSM parameter space, in the phenomenologically allowed region, and for SUSY parameters below $`1\text{ TeV}`$. Needless to say, not all of the maxima can be simultaneously attained as they are obtained for different values of the parameters. Perhaps the most noticeable result is that the decay into the lightest MSSM Higgs boson ($`tch^0`$) is the one that can be maximally enhanced, reaching values of order $`B(tch^0)10^4`$ that stay fairly stable all over the parameter space, and in particular for almost all the range of allowed Higgs boson masses in the MSSM.
For the sake of comparison, in Fig. 3 we show the maximized ratio for the competing decay $`tcg`$ as a function of the intergenerational mixing parameter between the second and the third generation, $`\delta _{23}`$ (4). We see that it never really reaches the critical value $`10^5,`$ which can be considered as the visible threshold for the next generation of high luminosity colliders. To assess the discovery reach of the FCNC top quark decays for these future accelerators we take as a guide the estimations that have been made for gauge boson final states . Assuming that all the FCNC decays $`tcX`$ ($`X=V,h`$) can be treated similarly, we roughly estimate the following sensitivities for $`100fb^1`$ of integrated luminosity:
$$\mathrm{๐๐๐}:B\text{ }\stackrel{>}{}\text{ }5\times 10^5;\mathrm{๐๐}:B\text{ }\stackrel{>}{}\text{ }5\times 10^4;\mathrm{๐๐๐๐๐}:B\text{ }\stackrel{>}{}\text{ }5\times 10^3.$$
Although the LHC seems to be the most sensitive machine to this kind of physics (due to its highest luminosity) the LC is also very competitive due to the cleanness of its environment which should allow a much more efficient isolation of the rare events. The upgraded Tevatron, unfortunately, looks not so promising in this respect.
To better assess the realistic possibilities for detecting the most serious FCNC top quark decay candidates, $`tch`$ and $`tcg`$, we remark that around the loci of maximal rates in parameter space the following situation is achieved
$$5\times 10^6\text{ }\stackrel{<}{}\text{ }B(tcg)_{\mathrm{max}}<B(tch^0)_{\mathrm{max}}\text{ }\stackrel{<}{}\text{ }5\times 10^4.$$
(6)
In both types of decays the dominant effects come from SUSY-QCD. However, it should not be undervalued the fact that the maximum electroweak rates for $`tch`$ can reach the $`10^6`$ level. Last but not least, we stress once again that the largest FCNC rate both from SUSY-QCD and SUSY-EW is precisely that of the lightest CP-even state ($`tch^0`$), which is the only Higgs channel that is phase-space available across the whole MSSM parameter space.
### 2.4 Two-body decays into $`R`$-odd particles
In principle there exist three possible two-body decays of the top quark into $`R`$-odd particles: $`t\stackrel{~}{b}\chi ^+`$, $`t\stackrel{~}{t}\stackrel{~}{g}`$ and $`t\stackrel{~}{t}\chi ^0`$. These decays were reviewed in . From the latest combined analysis of the four LEP experiments a lower bound on the squark and chargino masses between $`80\text{ GeV}`$ and $`90\text{ GeV}`$ is obtained . The exact bound depends on the assumptions of the analysis. The chargino channel is then highly disfavoured, and might be already closed at the end of the present LEP run. The light gluino window ($`m_{\stackrel{~}{g}}\text{ }\stackrel{<}{}\text{ }5\text{ GeV}`$) still exists, but its importance is everyday more marginal . Otherwise the direct limits from the Tevatron $`m_{\stackrel{~}{g}}\text{ }\stackrel{>}{}\text{ }200\text{ GeV}`$ apply, and the gluino decay channel is completely ruled out. So we are left with the neutralino decay channel as the most interesting decay. Note that light top-squarks are a natural feature in the MSSM due to the large top quark Yukawa coupling, and the presence of large mixing in the stop sector. A light top-squark (with small $`\mathrm{tan}\beta `$) also leads to an enhancement of $`R_b`$, pushing it closer to the measured value . Under the conditions of light stop and low $`\mathrm{tan}\beta `$ the branching ratio for the top quark decay into a neutralino and a stop can be at the 20% to 30% level. The strong sector one-loop radiative corrections to this partial decay width have been computed in . A key feature of the radiative corrections in which $`R`$-odd and $`R`$-even particles are both involved in external legs is that it is no longer possible to separate between standard and SUSY corrections, since neither of these subsets is finite by itself, so gluon and gluino loop contributions must be added up in order to obtain a finite result. As a consequence of that there appear terms which present non-decoupling, namely corrections proportional to $`\alpha _s\mathrm{log}(m_{\stackrel{~}{g}}/m_{\stackrel{~}{t}_a})`$.
The on-shell renormalization procedure must again deal with non-standard counterterms. In this case a counterterm for the stop mixing angle, $`\delta \theta _t`$, is necessary, even in the case of vanishing angle; this can be fixed by the condition that $`\delta \theta _t`$ cancels the one-loop mixing two-point function between the two stops at $`m_{\stackrel{~}{t}_2}`$ .
In Fig. 4a we see the tree-level prediction for the various branching ratios $`B(t\stackrel{~}{t}\chi _\alpha ^0)`$ as a function of the lightest stop mass. We see that large values of this branching ratio can be obtained. In Fig. 4b the radiative QCD corrections to each of the decays are plotted. Their absolute value lies in the range $`212\%`$, for the small values of the gluino mass ($`m_{\stackrel{~}{g}}180\text{ GeV}`$) used in this figure. For larger gluino masses the corrections are negative, and in the range $`m_{\stackrel{~}{g}}=15\text{ TeV}`$ they grow from $`12\%`$ to $`22\%`$ .
### 2.5 Three-particle decays
In order to have a consistent description of the top quark decay width at the order $`G_F\alpha `$ it is necessary to account for the three-body decays of the top quark as well. In Ref. all the possible three-body decays of the top quark were investigated.<sup>6</sup><sup>6</sup>6See also Ref. for a recent analysis of some of these modes, including $`R`$-parity violating decays. The aim in that work was to concentrate the analysis in regions of the parameter space in which the corresponding two-body decays were phase-space closed. With the latest bounds on SUSY particle masses some of these decays turn out to be phase-space closed. Here we briefly review those that still are kinematically allowed.
$`tb\tau ^+\nu _\tau `$: In the MSSM there exists, in addition to the gauge boson mediated channel, also the charged Higgs boson one. Of course, if the charged Higgs is lighter than the top quark then it would proceed through the two-body decay $`tH^+b`$ analyzed in section 2.2. But if the charged Higgs is heavier than the top quark, then the additional contribution to the $`\tau ^+\nu _\tau `$ final state from the $`H^+`$ mediated diagram is in the range of 1-3% for $`M_{H^\pm }<200\text{ GeV}`$ and large $`\mathrm{tan}\beta >40`$; $`tbW^+h^0`$: Its decay width can only reach at most $`\mathrm{\Gamma }/\mathrm{\Gamma }_{SM}^0\text{ }\stackrel{<}{}\text{ }3\times 10^4`$.
Next we just quote the maximum branching ratios for some decays into $`R`$-odd particles. $`tb\chi ^0\chi ^+`$: $`\mathrm{\Gamma }/\mathrm{\Gamma }_{SM}^0\text{ }\stackrel{<}{}\text{ }0.006`$; $`t\stackrel{~}{b}\tau ^+\stackrel{~}{\nu }_\tau `$: $`\mathrm{\Gamma }/\mathrm{\Gamma }_{SM}^0\text{ }\stackrel{<}{}\text{ }1\%`$; $`tb\stackrel{~}{g}\chi ^+`$: We remark that this decay could only be possible in the light gluino scenario, but even in this case $`\mathrm{\Gamma }/\mathrm{\Gamma }_{SM}^0\text{ }\stackrel{<}{}\text{ }4\%`$. Let us comment now a bit on the decay $`tb\stackrel{~}{\tau }\stackrel{~}{\nu }_\tau `$. This is an interesting process since, although it contains two $`R`$-odd particles in the final state, both of them are sleptons, which therefore evade the limits from hadron colliders, and being from the third generation they are expected to be light. Two Feynman diagrams contribute to the amplitude of this process, one of them involves the exchange of a charged Higgs particle. Part of the couplings between $`H^+`$ and the $`\stackrel{~}{\tau }\stackrel{~}{\nu }_\tau `$ state involve the combination $`A_\tau \mathrm{tan}\beta +\mu `$, which can naturally reach large values. It turns out that if the decay channel $`H^+\stackrel{~}{\tau }\stackrel{~}{\nu }_\tau `$ is closed (i.e. $`M_{H^\pm }<m_{\stackrel{~}{\tau }}+m_{\stackrel{~}{\nu }_\tau }`$) the value if this partial decay width is enhanced. The maximum value of the partial decay width is obtained for $`M_{H^\pm }`$ slightly below the sum of the slepton masses. With this condition, together with large $`\mathrm{tan}\beta \text{ }\stackrel{>}{}\text{ }40`$, and $`|\mu |\text{ }\stackrel{>}{}\text{ }100\text{ GeV}`$ we find that the values for the corresponding branching ratio can easily be of $`50\%`$ . Of course, this kind of precise alignment may be not the most natural expectation in the MSSM, but even if the parameters are not โoptimizedโ for a maximum branching ratio, there exist regions (in the large $`\mathrm{tan}\beta `$ scenario) where it can be relevant, say above the $`10\%`$ level. One of the possible signals for this decay channel includes a $`\tau ^{}`$, i.e. a โwrong signโ lepton, which could help in identifying this decay chain. See Ref. for a complete compilation of possible distinctive signals.
## 3 Top-quark pair production in $`e^+e^{}t\overline{t}`$
Besides the possibility of direct production of SUSY particles at sufficiently high energies, accurate investigations of the production of standard fermion pairs in $`e^+e^{}`$ annihilation offers the indirect search for virtual SUSY particles through quantum effects in terms of loop corrections. In particular the top quark, owing to its large mass, is an ideal probe of all those virtual effects that grow with the fermion mass scale, such as Yukawa interactions and CP violation. The MSSM with complex parameters induces CP-violating observables already at the one-loop level; they are described below after a brief discussion of the generic supersymmetric loop contributions to the CP-conserving cross section for $`t\overline{t}`$ production.
### 3.1 Loop effects in the CP-conserving MSSM
The process of $`t\overline{t}`$ production is in lowest order described by photon and $`Z`$-boson exchange. The cross section can be written as follows
$`\sigma ^{(0)}(e^+e^{}t\overline{t})={\displaystyle \frac{\beta }{4\pi s}}\left[{\displaystyle \frac{3\beta ^2}{2}}\sigma _V(s)+\beta ^2\sigma _A(s)\right]`$ (7)
with
$`s=(p_e^{}+p_{e^+})^2,\beta =\sqrt{1{\displaystyle \frac{4m_t^2}{s}}},`$ (8)
and
$`\sigma _V`$ $`=`$ $`Q_e^2Q_t^2e(s)^4+\mathrm{\hspace{0.17em}2}v_ev_tQ_eQ_te(s)^2\chi (s)+(v_e^2+a_e^2)v_t^2\chi (s)^2,`$
$`\sigma _A`$ $`=`$ $`(v_e^2+a_e^2)a_t^2\chi (s)^2.`$ (9)
This expression is an effective Born approximation, which incorporates the effective (running) electromagnetic charge containing the photon vacuum polarization (real part)
$`e(s)^2={\displaystyle \frac{4\pi \alpha }{1\mathrm{\Delta }\alpha (s)}},`$ (10)
the $`Z`$ propagator, together with the overall normalization factor of the neutral-current couplings in terms of the Fermi constant $`G_\mu `$,
$`\chi (s)=(G_\mu M_Z^2\sqrt{2})^2{\displaystyle \frac{s}{sM_Z^2}},`$ (11)
and the vector and axial-vector coupling constants for $`f=e,t`$,
$`v_f=I_3^f2Q_f\mathrm{sin}^2\theta _W,a_f=I_3^f.`$ (12)
The complete electroweak one-loop corrections were calculated for the case of the Standard Model , including also the hard-photon QED corrections . In a more recent study , a complete one-loop calculation was performed for the MSSM electroweak corrections and the non-standard part of the QCD corrections (SUSY-QCD corrections) to $`e^+e^{}t\overline{t}`$. The virtual Higgs contributions have also been derived in a general 2-Higgs-doublet model . For a large mass of the $`A^0`$ boson in the MSSM Higgs sector, the heavy particles $`H^0,A^0,H^\pm `$ decouple and the light $`h^0`$ scalar behaves like the standard Higgs boson. In this so-called decoupling limit, the only non-standard virtual SUSY effects arise from the genuine supersymmetric particles.
The one-loop contribution to the $`S`$-matrix element contains the $`\gamma `$ and $`Z`$ self-energies, the $`\gamma `$ and $`Z`$ vertex corrections together with the external wave function renormalization, and the box diagrams. Since Higgs-boson couplings to the initial state $`e^+,e^{}`$ are negligible, we only have to consider the standard box graphs with $`Z`$ and $`W^\pm `$ exchange and the SUSY box graphs with neutralino and chargino exchange.
The complete set of vertex corrections comprises the QED corrections with virtual photons and the QCD corrections with virtual gluons. They need real photon and gluon bremsstrahlung for a infrared-finite result. The gauge-invariant subclasses of โstandard QEDโ and โstandard QCDโ corrections are identical to those in the Standard Model and are available in the literature . In the meantime, also a combined treatment of the QCD and the electroweak radiative corrections in the Standard Model has been proposed in order to account for the dominant terms from both sources. In our context, we are interested in deviations from the Standard Model induced by non-standard virtual particles. We therefore concentrate our discussion on the set of model-dependent and IR-finite virtual corrections, without the diagrams involving virtual photons and gluons. The supersymmetric part of the QCD corrections, arising from virtual gluinos, is included as part of the final-state vertex correction.
The effect of these IR-finite loop-corrections is illustrated in Figure 5, where the energy dependence of lowest-order cross section (7) is shown together with the Standard Model (SM) and the MSSM one-loop prediction. The differences between the SM and the MSSM depend on the values of the SUSY-breaking parameters and the Higgsino mass parameter $`\mu `$. In case of a direct detection of SUSY particles, these parameters can, at least partially, be measured and the top production cross section can provide an independent test of the MSSM. In a first phase of a Linear Collider operation, the mass of eventually light SUSY particles, such as a scalar top $`\stackrel{~}{t}_1`$ or a chargino $`\chi _1^\pm `$ may be determined sufficiently well, while the heavier particles may appear as less accessible. A precise measurement of the $`t\overline{t}`$ cross section is therefore useful to probe the heavier part of the model through the quantum contributions. To make this more quantitative, we assume a scenario with a relatively light stop and chargino, and analyze the dependence of the loop-contributions on the residual parameters of the MSSM, which determine the heavier part of the mass spectrum. In order to exhibit the deviations from the SM, we introduce the quantity
$`\mathrm{\Delta }={\displaystyle \frac{\sigma ^{\mathrm{MSSM}}(s)\sigma ^{\mathrm{SM}}(s)}{\sigma ^{(0)}(s)}},`$ (13)
which gives the difference between the models normalized to the Born cross section (7). The results are displayed in Figure 6. For simplicity, we have assumed a common mass scale $`M_S`$ and a common $`A`$ parameter for the diagonal and non-diagonal entries in the sfermion mass-square matrices of all generations (as well as for $`L`$ and $`R`$ chiralities). The free parameters in Figure 6 are varied in accordance with a set of constraints which ensure consistency with the present electroweak precision data and bounds from direct Higgs searches. The effects are in the per-cent range, and show quite some sensitivity to the model parameters. Large values for $`\mathrm{tan}\beta `$ induce bigger loop effects, owing to the enhanced $`b`$ quark/squark Yukawa couplings.
As another possible scenario, we consider the situation that SUSY particles are not directly detected at a 500 GeV collider. The corresponding virtual SUSY effects, compatible with these bounds and with the constraints from the electroweak precision data, are shown in Figure 7. At least for the large-$`\mathrm{tan}\beta `$ regime the loop contributions are at the level of 1โ2 per cent.
### 3.2 CP violation in the MSSM
#### 3.2.1 The MSSM with complex couplings
The following parameters of the MSSM with preserved $`R`$-parity may take complex values: the Yukawa couplings, the $`\mu `$ parameter and the soft-breaking parameters $`m_{12}^2`$ (in the Higgs potential), $`M_1`$, $`M_2`$, $`M_3`$ (gaugino mass terms) and $`A`$ (trilinear terms). But not all of the complex phases introduced are physical since several of them can be absorbed by redefinitions of the fields.
We assume the GUT relation between the $`M_i`$, so they have one common phase. The remaining phases are chosen to be those of $`\mu `$, the $`A`$ parameters and, for three generations, one phase for all the Yukawa couplings, the $`\delta _{\mathrm{CKM}}`$. Analogous to the Standard Model case, the Yukawa couplings can be changed by redefinitions of the quark superfields in such a way that there remains only one phase for three generations.
Furthermore, the $`_{\mathrm{MSSM}}`$ is invariant under two U(1) transformations, the Peccei-Quinn symmetry and the $`R`$-symmetry, that do not only transform the fields but also the parameters of the MSSM . With them one can easily show that the physical predictions only depend on the absolute values of the parameters and the phases
$`\varphi _A\mathrm{arg}(AM_i^{}),\varphi _B\mathrm{arg}(\mu Am_{12}^2).`$ (14)
It is then possible to choose a set where only the $`\mu `$ and the $`A`$ parameters are complex. We take $`\phi _\mu =\mathrm{arg}(\mu )`$ and the phase of $`A`$ is traded for the phase $`\phi _{\stackrel{~}{f}}`$ of the off-diagonal term in the corresponding sfermion mixing matrix: $`m_{LR}^fA_f\mu ^{}\{\mathrm{cot},\mathrm{tan}\}\beta `$, for {up, down}-type fermions, respectively. Relaxing universality for the soft-breaking terms, every $`A_f`$ has a different phase.
In comparison with the MSSM with real couplings, the new physical phases affect the mass spectrum of charginos and neutralinos, modify the tree level couplings and notably produce CP-violating effects at the one-loop level that we analyze in the following.
#### 3.2.2 Electric dipole form factors
The most general Lorentz structure for the vertex $`Vff`$ in the momentum space is
$`\mathrm{\Gamma }_\mu ^{Vff}`$ $`=`$ $`\mathrm{i}[\gamma _\mu (f_\mathrm{V}f_\mathrm{A}\gamma _5)+(q\overline{q})_\mu (f_\mathrm{M}+\mathrm{i}f_\mathrm{E}\gamma _5)+p_\mu (\mathrm{i}f_\mathrm{S}+f_\mathrm{P}\gamma _5)`$ (15)
$`+(q\overline{q})^\nu \sigma _{\mu \nu }(f_{\mathrm{TS}}+\mathrm{i}f_{\mathrm{TP}}\gamma _5)+p^\nu \sigma _{\mu \nu }(\mathrm{i}f_{\mathrm{TM}}+f_{\mathrm{TE}}\gamma _5)],`$
where $`q`$ and $`\overline{q}`$ are the outgoing momenta of the fermions and $`p=(q+\overline{q})`$ is the total incoming momentum of the neutral vector boson $`V`$. The form factors $`f_i`$ are functions of kinematical invariants and can be complex in general. Their real parts account for dispersive effects (CPT-even) whereas their imaginary parts are related to absorptive contributions. For on-shell fermions, making use of the Gordon identities, one can eliminate $`f_{\mathrm{TM}}`$, $`f_{\mathrm{TE}}`$, $`f_{\mathrm{TS}}`$ and $`f_{\mathrm{TP}}`$. The number of relevant form factors can be further reduced when $`V`$ is on shell, since then the condition $`p_\mu ฯต^\mu =0`$ cancels the contributions coming from $`f_\mathrm{S}`$ and $`f_\mathrm{P}`$. The form factors $`f_\mathrm{V}`$ and $`f_\mathrm{A}`$ are connected to the chirality conserving, CP-even sector. The anomalous magnetic and electric dipole form factors are defined, respectively, by
AMDFF $``$ $`a_f^V(s)2m_ff_\mathrm{M}(s),`$ (16)
EDFF $``$ $`d_f^V(s)ef_\mathrm{E}(s).`$ (17)
They are related to chirality-flipping operators of dimension larger than four. In a renormalizable theory they can receive contributions exclusively by quantum corrections. Unlike the MDFFs, the EDFFs are connected to the CP-odd sector and therefore constitute a source of CP violation. The electromagnetic and weak dipole moments are physical and gauge invariant quantities corresponding to the values of the dipole form factors at $`s=(q+\overline{q})^2=M_V^2`$ for $`V=\gamma ,Z`$, respectively.
All the possible one-loop contributions to the $`a_f^V(s)`$ and $`d_f^V(s)`$ form factors can be classified in terms of the six classes of triangle diagrams. The vertices involved are labeled by generic couplings according to a general interaction Lagrangian. Every class of diagrams has been calculated analytically and expressed, in the โt Hooft-Feynman gauge, in terms of masses, generic (complex) couplings and one-loop 3-point integrals in Refs. .
We concentrate here on the CP-violating electric dipoles of the top quark. Unlike the SM, where the EDFFs arise first at three loops , in the MSSM they receive contributions already at the one-loop level . On the other hand, since the top quark is too heavy to be pair-produced by $`Z`$ decays, and has a too short life to study its electromagnetic static properties, the electric $`d_t^\gamma (s)`$ and weak-electric $`d_t^Z(s)`$ form factors, $`s>4m_t^2`$, are not physical. In fact, they do not contribute alone to CP-violating observables, as it will be show in next Section. Nevertheless they happen to be gauge invariant quantities at one loop.
The size of the electric dipole form factors of the top quark is illustrated in Fig. 8 where the different contributions from charginos, neutralinos and gluinos are also displayed. Neutralinos and gluinos appear with stops in the loops, and charginos with sbottoms. The Higgs sector does not play any role. The results depend on a number of MSSM parameters: $`\mathrm{tan}\beta `$, $`M_2`$, $`|\mu |`$, $`m_{LR}^t`$, $`m_{LR}^b`$ and $`m_{\stackrel{~}{q}}`$ (the common squark mass parameter); and three physical CP-violating phases, $`\phi _\mu `$, $`\phi _{\stackrel{~}{t}}`$ and $`\phi _{\stackrel{~}{b}}`$. They are expressed in $`t`$ magnetons $`\mu _te/2m_t=5.64\times 10^{17}e`$ cm. The values shown in the plots are for the set of inputs:
$`\mathrm{tan}\beta =1.6`$
$`M_2=|\mu |=m_{\stackrel{~}{q}}=|m_{LR}^t|=|m_{LR}^b|=200\text{ GeV}`$
$`\text{Reference Set }\mathrm{\#}1:`$ $`\phi _\mu =\phi _{\stackrel{~}{t}}=\phi _{\stackrel{~}{b}}=\pi /2.`$ (18)
This set of parameters is plausible and enhances the electric dipole form factors in the vicinity of $`\sqrt{s}=500`$ GeV, due to threshold effects (spikes in Fig. 8). The phases are chosen maximal and with such a sign that the typically larger components (charginos and gluinos) add up constructively. The low $`\mathrm{tan}\beta `$ scenario has been chosen because both chargino and gluino contributions are then larger. Of course, the electric dipoles vanish for zero phases. Due to the decoupling of the supersymmetric sector, the dipoles tend to vanish for increasing values of the mass parameters. A detailed study can be found in .
#### 3.2.3 CP-odd observables
Not only vertex- but also box-diagrams correct the tree level process to one loop. Accordingly, any CP-odd observable will depend on the CP-violating effects due to vertex corrections included in the electric and weak-electric dipole form factors as well as on CP-violating box contributions. In the former there appear either sneutrinos, neutralinos and stops or selectrons, charginos and sbottoms.
To probe CP violation, consider the pair-production of polarized $`t`$ quarks $`e^+(๐ฉ_+)+e^{}(๐ฉ_{})t(๐ค_+,๐ฌ_\mathrm{๐})+\overline{t}(๐ค_{},๐ฌ_\mathrm{๐})`$. The decay channels labeled by $`a`$ and $`c`$ act as spin analyzers in $`t+\overline{t}a(๐ช_+)+\overline{c}(๐ช_{})+X`$. The momenta and polarization vectors in the overall c.m.s. transform under CP and CPT as follows:<sup>7</sup><sup>7</sup>7T means reflection of spins and momenta.
$`\begin{array}{cc}\hfill \text{CP}:& ๐ฉ_\pm ๐ฉ_{}=๐ฉ_\pm \hfill \\ & ๐ค_\pm ๐ค_{}=๐ค_\pm \hfill \\ & ๐ช_\pm ๐ช_{}\hfill \\ & ๐ฌ_\mathrm{๐}๐ฌ_\mathrm{๐}\hfill \end{array}`$ $`\begin{array}{cc}\hfill \text{CPT}:& ๐ฉ_\pm ๐ฉ_{}=๐ฉ_\pm \hfill \\ & ๐ค_\pm ๐ค_{}=๐ค_\pm \hfill \\ & ๐ช_\pm ๐ช_{}\hfill \\ & ๐ฌ_\mathrm{๐}๐ฌ_\mathrm{๐}\hfill \end{array}`$ (27)
From the unit momentum of one of the $`t`$ quarks in the c.m.s. (say $`\widehat{๐ค}_+`$) and their polarizations ($`๐ฌ_\mathrm{๐}`$, $`๐ฌ_\mathrm{๐}`$) a basis of linearly independent CP-odd spin observables can be constructed . The spin observables are related to more realistic (directly measurable) momentum observables based on the momenta of the top decay products . The polarizations can be analyzed through the angular correlation of the weak decay products, both in the nonleptonic and in the semileptonic channels:
$`t(๐ค_+)b(๐ช_๐)X_{\mathrm{had}}(๐ช_๐),`$ (28)
$`t(๐ค_+)b(๐ช_๐)\mathrm{}^+(๐ช_+)\nu _{\mathrm{}}(\mathrm{}=e,\mu ,\tau )`$ (29)
and the charged conjugated ones.
#### 3.2.4 Spin observables
A list of CP-odd spin observables classified according to their CPT properties is shown in Table 1. Their expectation values as a function of $`s`$ and the scattering angle of the $`t`$ quark in the overall c.m. frame are given by
$`๐ช_{\mathrm{๐๐}}`$ $`=`$ $`{\displaystyle \frac{1}{2\mathrm{d}\sigma }}\left[{\displaystyle \underset{๐ฌ_\mathrm{๐}^{},๐ฌ_\mathrm{๐}^{}=\pm ๐,\pm ๐}{}}+{\displaystyle \underset{๐ฌ_\mathrm{๐}^{},๐ฌ_\mathrm{๐}^{}=\pm ๐,\pm ๐}{}}\right]\mathrm{d}\sigma (๐ฌ_\mathrm{๐}^{},๐ฌ_\mathrm{๐}^{})๐ช,`$ (30)
$`\mathrm{d}\sigma `$ $`=`$ $`{\displaystyle \underset{\pm ๐ฌ_\mathrm{๐}^{},\pm ๐ฌ_\mathrm{๐}^{}}{}}\mathrm{d}\sigma (๐ฌ_\mathrm{๐}^{},๐ฌ_\mathrm{๐}^{}).`$ (31)
They are displayed in Fig. 10, for two sets of MSSM parameters, Set #1 of (18) and Set #2 with same moduli but different phases:
$`\text{Reference Set }\mathrm{\#}2:`$ $`\phi _\mu =\phi _{\stackrel{~}{t}}=\phi _{\stackrel{~}{b}}=\pi /2.`$ (32)
The directions of polarization of $`t`$ and $`\overline{t}`$ (a and b) are taken normal to the scattering plane (N), transversal (T) or longitudinal (L). They are taken either parallel ($``$) or antiparallel ($``$) to the axes defined by $`\widehat{z}=๐ค_+`$, $`\widehat{y}=๐ค_+\times ๐ฉ_+/|๐ค_+\times ๐ฉ_+|`$ and $`\widehat{x}=\widehat{y}\times \widehat{z}`$. The spin vectors $`๐ฌ_\mathrm{๐}^{}`$, $`๐ฌ_\mathrm{๐}^{}`$ are defined in the $`t`$, $`\overline{t}`$ rest frames, respectively. The statistical significance (number of standard deviations) of the CP-violation signal is given by
$`N_{SD}=|r|\sqrt{N},r๐ช/\sqrt{๐ช^2}`$ (33)
where $`N`$ is the number of observed events.
We compare the result when only the self energies and vertex corrections are included (left column) with the complete one-loop calculation (right column). The shape of the observables as a function of the $`t`$ polar angle is also quite different when the CP violating box contributions are taken into account (Fig. 10). This illustrates that the dipole form factors of the $`t`$ quark are not sufficient to parameterize observable CP-violating effects and the predictions can be wrong by far.
#### 3.2.5 Momentum observables
Consider now the decay channels labeled by $`a`$ and $`c`$ acting as spin analyzers in $`t+\overline{t}a(๐ช_+)+\overline{c}(๐ช_{})+X`$. We ignore here possible CP violation in the top-quark decays, that may occur due to supersymmetric one-loop corrections . The expectation value of a realistic CP-odd observable is given by the average over the phase space of the final state particles,
$`๐ช_{ac}={\displaystyle \frac{1}{2}}\left[๐ช_{a\overline{c}}+๐ช_{c\overline{a}}\right]={\displaystyle \frac{1}{2\sigma _{ac}}}{\displaystyle \left[\mathrm{d}\sigma _{a\overline{c}}+\mathrm{d}\sigma _{c\overline{a}}\right]๐ช},`$ (34)
where both the process ($`a\overline{c}`$) and its CP conjugate ($`c\overline{a}`$) are included and
$`\sigma _{ac}={\displaystyle d\sigma _{a\overline{c}}}={\displaystyle d\sigma _{c\overline{a}}},`$ (35)
The differential cross section for $`t`$-pair production and decay is evaluated for every channel using the narrow width approximation. As $`m_t>M_W+m_b`$, the $`t`$ quark decays proceed predominantly through $`Wb`$. Within the SM the angular distribution of the charged lepton is a much better spin analyzer of the $`t`$ quark than that of the $`b`$ quark or the $`W`$ boson arising from semileptonic or nonleptonic $`t`$ decays . The dimensionless observables are easier to measure, for instance the scalar CP-odd observables :
$`\widehat{A}_1`$ $``$ $`\widehat{๐ฉ}_+{\displaystyle \frac{\widehat{๐ช}_+\times \widehat{๐ช}_{}}{|\widehat{๐ช}_+\times \widehat{๐ช}_{}|}}\text{[CPT-even]}`$ (36)
$`\widehat{A}_2`$ $``$ $`\widehat{๐ฉ}_+(\widehat{๐ช}_++\widehat{๐ช}_{})\text{[CPT-odd]}`$ (37)
or the CP-odd traceless tensors :
$`\widehat{T}_{ij}`$ $``$ $`(\widehat{๐ช}_+\widehat{๐ช}_{})_i{\displaystyle \frac{(\widehat{๐ช}_+\times \widehat{๐ช}_{})_j}{|\widehat{๐ช}_+\times \widehat{๐ช}_{}|}}+(ij)\text{[CPT-even]}`$ (38)
$`\widehat{Q}_{ij}`$ $``$ $`(\widehat{๐ช}_++\widehat{๐ช}_{})_i(\widehat{๐ช}_+\widehat{๐ช}_{})_j+(ij)\text{[CPT-odd]}`$ (39)
The reconstruction of the $`t`$ frame is not necessary for the momentum observables. The observables $`\widehat{A}_2`$ and $`\widehat{Q}_{ij}`$ do not involve angular correlations as they could be measured considering separate samples of events in the reactions $`e^+e^{}aX`$ and $`e^+e^{}\overline{a}X`$. Nevertheless it is convenient to treat them in an event-by-event basis .
In Table 2 the ratio $`r`$ is shown for three different decay channels at $`\sqrt{s}=500`$ GeV. The CP-odd observables involve the momenta of the decay products analyzing $`t`$ and $`\overline{t}`$ polarizations in the laboratory frame. The leptonic decay channels are the best $`t`$ spin analyzers but the number of leptonic events is also smaller. The Reference Set $`\mathrm{\#}2`$ has been chosen. As expected, the dipole contributions (left columns) to the CPT even observables are very small for this choice of MSSM parameters but the actual expectation values (right columns) are larger.
#### 3.2.6 Polarized beams
The previous results were obtained for unpolarized electron and positron beams. Let $`P_\pm `$ be the degree of longitudinal polarization of the initial $`e^\pm `$. The differential cross section reads now
$`\mathrm{d}\sigma ={\displaystyle \frac{1}{4}}\left[(1+\mathrm{P}_+)(1+\mathrm{P}_{})\mathrm{d}\sigma _R+(1\mathrm{P}_+)(1\mathrm{P}_{})\mathrm{d}\sigma _L\right],`$ (40)
where $`\sigma _{R/L}`$ corresponds to the cross section for electrons and positrons with equal right/left-handed helicity. Chirality conservation suppresses opposite helicities. Table 3 summarizes some extreme cases. If both beams are fully polarized, $`\mathrm{P}_+=\mathrm{P}_{}=\pm 1`$, the ratio $`r`$ is the same as for ($`\mathrm{P}_\pm =0`$, $`\mathrm{P}_{}=\pm 1`$), respectively, but the cross sections are twice as much (40), which results in a higher statistical significance of the CP signal. From comparison of Table 2 (right columns) with Table 3 is clear that left-handed polarized beams enhance the sensitivity to CP-violating effects.
Concerning the experimental reach of this analysis, the statistical significance of the signal of CP violation is given by $`N_{SD}=|r|\sqrt{N}`$ with $`N=ฯต\sigma _{t\overline{t}}\text{BR}(ta)\text{BR}(\overline{t}\overline{c})`$ where $`ฯต`$ is the detection efficiency and $``$ the integrated luminosity of the collider. The branching ratios of the $`t`$ decays are BR $`1`$ for the $`b`$ channel and BR $`0.22`$ for the leptonic channels ($`\mathrm{}=e,\mu `$). At $`\sqrt{s}=500`$ GeV the total cross section for $`t`$-pair production is $`\sigma _{t\overline{t}}0.5`$ pb. Assuming a LC integrated luminosity $`500`$ fb<sup>-1</sup> and a perfect detection efficiency, one gets $`\sqrt{N}500,235,110`$ for the channels $`bb`$, $`\mathrm{}b`$, $`\mathrm{}\mathrm{}`$, respectively. With these statistics, values of $`|r|2\times 10^3,4\times 10^3,9\times 10^3`$ for the channels above would be necessary to achieve a 1 s.d. effect. Such values can be hardly reached in the context of the MSSM for the given luminosity, even for polarized beams, as Tables 2 and 3 show.
## 4 Conclusions
In summary, the top-quark is an excellent laboratory to test the MSSM. If SUSY is realized in nature around the electroweak scale, the top-quark phenomenology differs significantly from the SM one. The various top-quark observables can be well investigated at TESLA. The production cross-section receives radiative corrections that can be used to test the MSSM, or to derive bounds on the masses of the various sparticles. Also, in the MSSM there exist new sources of CP-violation effects, giving rise to specific CP-violating asymmetries in top-pair production, which, however, might be too small for being observable with the currently expected luminosity. The top-quark decay modes can differ significantly from the SM ones; of special importance is the $`tH^+b`$ decay mode. Even in the case that the new particles of the MSSM are heavier than the top-quark, their virtual quantum effects influence the top-quark standard decay mode. Moreover, the LC presents a great sensitivity to the possible rare decays. In the MSSM some of these decays are strongly enhanced with respect to the SM expectations, and could be detected.
## Acknowledgments
This work has been partially supported by the Deutsche Forschungsgemeinschaft, by CICYT under projects AEN99-0766 and AEN96-1672 and by Junta de Andalucรญa under project FQM-101. |
warning/0003/hep-ph0003213.html | ar5iv | text | # Photoproduction of Vector Mesons at Large Transfer
## Abstract
At forward angles, the cross-sections of photoproduction of vector mesons ($`\rho `$, $`\omega `$, and $`\varphi `$) are well accounted for by the exchange of the Pomeron at high energies, while contributions of $`t`$ channel exchange of Reggeons are significant at low energies. At large angles, the impact parameter becomes small enough to prevent their constituents to build up the exchanged Reggeons or Pomeron. Two gluon exchange appears to dominate above $`t1`$ GeV<sup>2</sup>, especially in the $`\varphi `$ channel.
Elastic photoproduction of vector mesons exhibits the same slow rise with the energy as the hadronic cross-sections (Fig. 1): The time during which the photon fluctuates into vector mesons is long enough to permit their interactions with the target nucleon. The exchange of the Pomeron accounts fairly well for this universal behavior, while at lower energies the exchange of a few Regge trajectories is necessary to reproduce the variation of the cross-sections . At large angles, the impact parameter is small enough to prevent two gluons (resp. two quarks) to build up the exchanged Pomeron (resp. Reggeons). This note investigates under which conditions the Pomeron and the Reggeons are resolved into their simplest constituents and how they couple to the constituent quarks in the vector meson and the nucleon.
Following the notations of Ref. , the amplitudes describing the exchange of the Pomeron and the $`f_2`$ meson can be combined in the compact form:
$`๐ฏ_P+๐ฏ_{f_2}=i{\displaystyle \frac{12\sqrt{6}m_Ve_qf_V\beta _0^2\mu _0^2F_1(t)}{(Q^2+m_V^2t)(2\mu _0^2+Q^2+m_V^2t)}}`$ (1)
$`\times [2pqฯต_Vฯต+2ฯต_Vpฯต(qP_V)2ฯต_Vqฯตp`$ (2)
$`+{\displaystyle \frac{Q^2+m_V^2t}{pq}}ฯต_Vpฯตp](\left({\displaystyle \frac{s}{s_0}}\right)^{\alpha _P(t)1}e^{\frac{1}{2}i\pi \alpha _P(t)}`$ (3)
$`+\kappa _{f_2}\left({\displaystyle \frac{s}{s_1}}\right)^{\alpha _{f_2}(t)1}{\displaystyle \frac{(1+e^{i\pi \alpha _{f_2}(t)})\pi \alpha _{f_2}^{}}{2\mathrm{sin}(\pi \alpha _{f_2}(t))\mathrm{\Gamma }(\alpha _{f_2}(t))}})`$ (4)
being $`p=(E_i,\stackrel{}{p_i})`$, $`q=(\omega ,\stackrel{}{k})`$ and $`P_V=(P_V^0,\stackrel{}{P_V})`$ the four momenta of the target nucleon, the incoming photon and the outgoing meson: $`2pq=sm^2+Q^2`$. The Regge trajectory of the Pomeron is $`\alpha _P(t)=1.08+0.25t`$, while the trajectory of the $`f_2`$ meson is $`\alpha _{f_2}(t)=0.55+\alpha _{f_2}^{}t`$, with $`\alpha _{f_2}^{}=0.7`$ GeV<sup>-2</sup>. The mass scales are $`s_0=4`$ GeV<sup>2</sup> and $`s_1=`$ 1 GeV<sup>2</sup>.
The model assumes that the Pomeron, or the $`f_2`$, couples to a single constituent quark in the vector meson and in the nucleon, as a $`C=+1`$ isoscalar photon. In the high energy limit, the trace over the matrices in the quark loops leads to the three first terms in the square bracket. The last term has been added to restore gauge invariance: it can be considered as a contact term. When the strength of the coupling of the Pomeron with a quark, $`\beta _0^24`$ GeV<sup>2</sup>, is fixed by the analysis of the nucleon-nucleon high energy scattering, the Pomeron exchange amplitude reproduces the magnitude of vector meson photoproduction cross sections in the HERA energy range, $`W100`$ GeV. Their ratio follows from the actual values of the effective charge $`e_q`$ of the quark, the mass $`m_V`$ and the radiative decay constant $`f_V`$ of the meson. In the Fermi Lab, CERN and SLAC energy range, $`W10`$ GeV and below, the Pomeron exchange contribution underestimates the measured cross sections. Here the exchange of a trajectory with an intercept close to $`0.5`$ is needed to provide a contribution which decreases slowly with energy. Since the $`\rho `$ or $`\omega `$ meson cannot be exchanged, the model includes a non-degenerate trajectory based on the $`f_2(1270)`$ meson, which is assumed to couple to the quark as the Pomeron, the strength of its coupling ($`\kappa _{f_2}=9`$) being adjusted in order to reproduce the $`\rho `$ channel cross section. It reproduces also the $`\omega `$ channel cross section, especially its natural parity exchange component (open squares at $`W=2`$ and $`3`$ GeV, in Fig. 1). In the $`\varphi `$ channel, the Pomeron exchange amplitude reproduces the trend of the cross-section down to threshold, slightly overestimating it (by 20%) at low energy. Here, the exchange of $`f_2^{}(1525)`$, which has a significant strangeness content, is preferred to the $`f_2(1270)`$ meson and brings the model close to the data when $`\kappa _{f_2^{}}=4.5`$.
Scalar and pseudo scalar meson exchanges contribute at lower energies accessible at SLAC and JLab. The spatial part of the $`\pi `$ exchange amplitude takes the form:
$`๐ฏ_\pi =i{\displaystyle \frac{eg_{V\pi \gamma }}{m_\pi }}g_0{\displaystyle \frac{\sqrt{(E_i+m)(E_f+m)}}{2m}}F_\pi (t)\left({\displaystyle \frac{s}{s_1}}\right)^{\alpha _\pi (t)}`$ (5)
$`{\displaystyle \frac{\pi \alpha _\pi ^{}e^{i\pi \alpha _\pi (t)}}{\mathrm{sin}(\pi \alpha _\pi (t))\mathrm{\Gamma }(\alpha _\pi (t)+1)}}\left(\chi _f\right|\stackrel{}{\sigma }[{\displaystyle \frac{\stackrel{}{p_i}}{E_i+m}}`$ (6)
$`{\displaystyle \frac{\stackrel{}{p_f}}{E_f+m}}]\left|\chi _i\right)[\stackrel{}{k}\times (ฯต_V^0\stackrel{}{P_V}P_V^0\stackrel{}{ฯต_V})\omega \stackrel{}{ฯต_V}\times \stackrel{}{P_V}]\stackrel{}{ฯต}`$ (7)
being $`(E_f,\stackrel{}{p_f})`$ the four momenta of the recoiling nucleon and $`\chi _i`$, $`\chi _f`$ the nucleon spinors. The degenerate Regge trajectory is $`\alpha _\pi (t)=(tm_\pi ^2)\alpha _\pi ^{}`$, with $`\alpha _\pi ^{}=0.7`$ GeV<sup>-2</sup>. The $`\pi N`$ coupling constant is $`g_0^2/4\pi =14.5`$. According to the low energy phenomenology, a monopole form factor $`F_\pi (t)=(\mathrm{\Lambda }_\pi ^2m_\pi ^2)/(\mathrm{\Lambda }_\pi ^2t)`$, with $`\mathrm{\Lambda }_\pi =1.1`$ GeV, takes into account the finite size of the $`\pi N`$ vertex.
The $`\pi `$ exchange amplitude dies quickly when the energy increases, due to the nearly vanishing intercept of the $`\pi `$ trajectory. It contributes dominantly in the $`\omega `$ channel (Fig. 1), due to the relative size of the radiative coupling constants: $`g_{\omega \pi \gamma }=0.334`$, $`g_{\rho \pi \gamma }=0.136`$, $`g_{\varphi \pi \gamma }=0.0301`$. In the $`\varphi `$ channel, $`\eta `$ exchange may also be considered , but due to the actual size of the coupling constants it contributes at the same level as the $`\pi `$ exchange and is negligible.
In the $`\rho `$ production channel, $`\sigma `$ exchange cannot be excluded near threshold. The amplitude follows the expression of Ref. , but with a degenerate Regge propagator. The radiative coupling constant has been adjusted in order to accommodate the highest values of the experimental cross section, but the spread of the data can also be accounted for by the $`f_2`$ exchange alone.
It is remarkable that such a simple model reproduces over two decades the energy variation of the cross section of the three channels, at the expense of only one parameter, the relative strength of the $`f_2`$ effective coupling. All the other inputs (Pomeron quark or $`\pi `$N couplings, radiative couplings, Regge trajectories, โฆ) are determined from other channels. As illustrated in Fig. 2, it naturally reproduces the forward angle cross-sections as well as the shrinking of the diffractive peak when the energy increases from the SLAC to the HERA domain (dotted lines).
However, it underestimates the experiment at large transfer $`t=(P_Vq)^2`$: above $`t=0.5`$ GeV<sup>2</sup> in the HERA energy range and $`t=1`$ GeV<sup>2</sup> in the JLab/SLAC energy range. Here, the impact parameter ($`b1/\sqrt{t}0.2`$ fm when $`t=1`$ GeV<sup>2</sup>) becomes comparable or smaller to the gluon correlation length ($`a0.2รท0.3`$ fm)โ the distance over which a gluon can propagate: two gluons can be exchanged between a quark in the vector meson and a quark in the proton, without being forced to recombine into a Pomeron. Larger momentum transfers also select configurations where the transverse distance between the two quarks, which must recombine into the vector meson, is small: each gluon can also couple to a different quark. As demonstrated in Ref. , the interference between these two mechanisms leads to a characteristic node, around $`t2รท2.5`$ GeV<sup>2</sup>, in the two gluon exchange cross-section (dashed lines). The color singlet nature of the $`q\overline{q}`$ pair imposes such a destructive interference. It is worth noting that such a cancellation is at the origin of color transparency : the weakening of the interaction between white objects of which the transverse size vanishes. On the same token, each gluon can also couple to a different quark inside the nucleon target, giving access to correlations between quarks in its ground state (full lines).
Under the same assumptions, and with the same notations, as in Ref. the two gluon amplitude takes the form:
$`๐ฏ_{2g}={\displaystyle \frac{6\sqrt{6}m_Ve_qf_V}{8\pi }}[2pqฯต_Vฯต+2ฯต_Vpฯต(qP_V)`$ (8)
$`2ฯต_Vqฯตp+{\displaystyle \frac{Q^2+m_V^2t}{pq}}ฯต_Vpฯตp]`$ (9)
$`{\displaystyle dl^2[4\pi \alpha _sD(l^2+\frac{1}{4}t)]^2[F_1(t)F_1(3l^2+\frac{t}{4})]}`$ (10)
$`\left({\displaystyle \frac{1}{Q^2+m_V^2t}}{\displaystyle \frac{1}{Q^2+m_V^24l^2}}\right)`$ (11)
where the integral runs over the transverse momentum $`l^2`$ of the exchanged gluons. It follows from the works of Refs. and is the same as the expression derived in Ref. , except that, instead of a pertubative propagator, a dressed gluon propagator is used:
$`\alpha _sD(l^2)={\displaystyle \frac{\beta _0}{\sqrt{\pi }\lambda _0}}\mathrm{exp}\left({\displaystyle \frac{l^2}{\lambda _0^2}}\right),`$ (12)
$`2\pi {\displaystyle _{\mathrm{}}^0}dl^2[\alpha _sD(l^2)]^2=\beta _0^2`$ (13)
The value of the range parameter $`\lambda _0^2=2.7`$ GeV<sup>2</sup> corresponds to a gluon correlation length $`a0.19`$ fm.
The gauge invariant expression in the first square bracket results from the trace in the parton loops, assuming that the constituent quarks of the vector meson are frozen. In Refs. , the two gluons were assumed to couple to the same quark of the nucleon, the structure of which being taken into account by the experimental value of the isoscalar form factor
$`F_1(t)={\displaystyle \left[\mathrm{d}\beta _j\mathrm{d}\stackrel{}{r_j}\right]\delta (\beta _j\stackrel{}{r_j})\delta (\beta _j1)}`$ (14)
$`|\mathrm{\Psi }(\beta _j,\stackrel{}{r_j})|^2e^{i\stackrel{}{\mathrm{\Delta }}\stackrel{}{r_k}}`$ (15)
where the nucleon wave function $`\mathrm{\Psi }`$ depends upon the transverse coordinate $`\stackrel{}{r_j}`$ and the fraction of the longitudinal momentum $`\beta _j`$ carried by the quark $`j`$, and where $`\stackrel{}{\mathrm{\Delta }}^2=t`$. When each gluon is allowed to couple to a different quark, the structure of the nucleon enters eq. 11 through the correlation function
$`G(\stackrel{}{k_a},\stackrel{}{k_b})={\displaystyle \left[\mathrm{d}\beta _j\mathrm{d}\stackrel{}{r_j}\right]\delta (\beta _j\stackrel{}{r_j})\delta (\beta _j1)}`$ (16)
$`|\mathrm{\Psi }(\beta _j,\stackrel{}{r_j})|^2e^{i\stackrel{}{k_a}\stackrel{}{r_k}+i\stackrel{}{k_b}\stackrel{}{r_l}}`$ (17)
Where $`\stackrel{}{k_a}=\stackrel{}{l}+\frac{1}{2}\stackrel{}{\mathrm{\Delta }}`$ and $`\stackrel{}{k_b}=\stackrel{}{l}\frac{1}{2}\stackrel{}{\mathrm{\Delta }}`$ are the momenta of the exchanged gluons. Under the assumption that the gluons couple to valence quarks equally sharing the longitudinal momentum, the nucleon wave function takes the form
$`\mathrm{\Psi }(\beta _j,\stackrel{}{r_j})\left({\displaystyle \delta (\beta _i\frac{1}{3})}\right)\mathrm{exp}\left({\displaystyle \frac{\stackrel{}{r_i}^2}{r_N^2}}\right)`$ (18)
and the correlation function can be expressed as the isosclar form factor at a different argument
$`G(\stackrel{}{k_a},\stackrel{}{k_b})=F_1(3l^2+{\displaystyle \frac{t}{4}})`$ (19)
Each of the two last terms in eq. 11 corresponds to the propagator of the quark which is off-shell in the vector meson loop, when the two gluons couple either to the same quark or to two different quarks.
Eq. 11 exhibits in a compact way the main features of the two gluon exchange mechanisms, without free parameters. The strength, $`\beta _0^2=4`$ GeV<sup>2</sup>, of the effective coupling of the gluon to the nucleon is determined by the analysis of the nucleon-nucleon scattering, the range parameter $`\lambda _0^2`$ of the gluon dressed propagator corresponds to a reasonable value of the gluon correlation length, and the correlated structure of the nucleon enters through the experimental values of its form factor. It provides us with a solid starting ground for looking for such a mechanism in experiments.
A recent experiment performed at HERA confirms this picture. As shown in Fig. 2, these data clearly favor two gluon exchange above $`t=0.5`$ GeV<sup>2</sup> but, due to the statistics, cannot distinguish between the two extreme two gluon curves. They differ above $`t=1`$ GeV<sup>2</sup> where a old experiment performed at SLAC , at lower energy, confirms that the mechanism which dominates at high momentum transfer is the coupling of each gluon to two different quarks both in the vector meson and in the nucleon.
However, quark exchange cannot be excluded in the $`\rho `$ production channel, in the SLAC/CEBAF energy range. Their contribution has been evaluated using saturating Regge trajectories ($`\alpha _M(\mathrm{})=1`$) in the $`\pi `$, $`f_2`$ and $`\sigma `$ meson exchange amplitudes, as explained in Ref. . The $`\pi `$ exchange process leads to a contribution too small by two orders of magnitude. The $`f_2`$ exchange leads to a vanishing contribution, due to the signature $`1+e^{i\pi \alpha _{f_2}(t)}`$ of its undegenerated trajectory. Only $`\sigma `$ exchange may lead to a contribution comparable to the data, but this is unconclusive due to the uncertainty on the amount of its contribution at low energy (cf. Fig. 1).
Due to its dominant $`s\overline{s}`$ component, quark exchange mechanisms are strongly suppressed in $`\varphi `$ photoproduction, making this channel the most suitable place to look for gluon exchanges. Fig. 3 compares model predictions to data recently obtained at HERA as well as previous data (see for references) in the JLab energy range. As in the $`\rho `$ channel, the Regge model reproduces the data up to $`t=0.5`$ GeV<sup>2</sup> in the HERA energy range and $`t=1`$ GeV<sup>2</sup> in the JLab energy range, and underestimate the data above. At high transfer the HERA data clearly prefer the two gluon exchange model, but are unable to distinguish between its two extreme versions. This is the goal of JLab experiment 93-031, which has been recently completed in the domain $`3E_\gamma 5.5`$ GeV, $`1t6`$ GeV<sup>2</sup>. It is remarkable that the preliminary results fall on the full curve in the bottom panel of Fig. 3. A detailed comparison with those data will be available in the experimental paper .
At backward angles (largest $`t`$ but smallest $`u`$) exchange of baryons in the u-channel may contribute in the JLab energy range. Only the exchange of the nucleon Regge trajectory is allowed in the $`\varphi `$ (as well as $`\omega `$) channel. The spatial part of the amplitude is as follows:
$`๐ฏ_N=i{\displaystyle \frac{e\mu _pg_V(1+\kappa _V)}{2m}}\left(\chi _f\left|\stackrel{}{\sigma }\stackrel{}{k}\times \stackrel{}{ฯต}\stackrel{}{\sigma }\stackrel{}{P_V}\times \stackrel{}{ฯต_V}\right|\chi _i\right)`$ (20)
$`\left({\displaystyle \frac{s}{s_1}}\right)^{\alpha _N(u)\frac{1}{2}}{\displaystyle \frac{\pi \alpha _N^{}(1e^{i\pi (\alpha _N(u)+\frac{1}{2})})}{2\mathrm{sin}(\pi (\alpha _N(u)+\frac{1}{2}))\mathrm{\Gamma }(\alpha _N(u)+\frac{1}{2})}}`$ (21)
where the non degenerate nucleon trajectory is $`\alpha _N(t)=0.37+\alpha _N^{}t`$, with $`\alpha _N^{}=0.98`$ GeV<sup>-2</sup>. This amplitude leads to a good accounting of the backward angle cross-section in the $`\omega `$ channel where all the coupling constants are fixed. In the $`\varphi `$ production channel, the dot-dashed curve in Fig. 3 results from the SU3 choice of the unknown $`\varphi `$NN coupling constant $`g_V(1+\kappa _V)`$. In the rho channel, the $`\mathrm{\Delta }`$ degenerate trajectory may be exchanged: it interferes with nucleon exchange and leads to a good account of the backward angle $`\rho `$ meson photoproduction (dot-dashed line in Fig. 2).
In the highest part of the JLab energy range, the $`u`$-channel exchange backward peak and the diffractive forward peak are far enough to leave room for hunting gluon exchange processes, and studying the structure of the proton. It remains to investigate whether effects, which are negligible at higher energies, play a role: real part of the amplitude, motion of the quarks in the vector meson loop , full energy dependence , etcโฆ
In summary, two gluon exchange mechanisms are good candidate to understand vector meson photoproduction at large $`t`$. The model which we have developed provides a good starting ground as it incorporates the basic features and its inputs are strongly constrained by other channels. It reproduces data recently obtained in two extreme energy domains: HERA and JLab. It provides us with a tool to investigate quark correlations in the nucleon ground state and to test more realistic wave functions.
Discussions with M. Diehl and M. Vanderhaeghen are greatly acknowledged. |
warning/0003/cond-mat0003340.html | ar5iv | text | # Collective Modes in Strongly Coupled Elecronic Bilayer Liquids
## Abstract
We present the first reliable calculation of the collective mode structure of a strongly coupled electronic bilayer. The calculation is based on a classical model through the $`3^{rd}`$ frequency-moment-sum-rule preserving Quasi Localized Charge Approximation, using the recently calculated Hypernetted Chain pair correlation functions. The spectrum shows an energy gap at $`k=0`$ and the absence of a previously conjectured dynamical instability.
Electronic bilayers exhibit a rich pattern of behavior, both on the static and on the dynamic level. While at high $`r_s`$ values ($`r_s>r_s^{crystal}`$), the bilayer is expected to crystallize (according to , $`r_s^{crystal}20`$), and at very low $`r_s`$ values the RPA description is largely sufficient, the most interesting behavior occurs in the $`1<r_s<r_s^{crystal}`$ domain, where the system is in the liquid state. This is the domain we focus on in this Letter.
The most remarkable feature on the static level is that the system exhibits a series of abrupt structural changes as the ratio of the interlayer distance $`d`$ and the 2D Wigner-Seitz radius $`a`$ is varied . These structural changes parallel the structural phase transitions in the solid phase , but at finite temperature they are also combined with entropy increasing substitutional order-disorder transitions where particles in layer $`1(2)`$ occupy positions appropriate for particles in layer $`2(1)`$: this is signalled by the two pair correlation functions (PCF) $`h_{11}(r)`$ and $`h_{12}(r)`$ becoming identical as $`d0`$.
The dynamical behavior, the collective mode structure in particular, has been studied by Swierkowski et al. , by Gold , by Zhang and Tzoar , Golden, Kalman and collaborators and by Moudgil et al. . (Some of these studies pertain to a superlattice \[infinite number of layers\], rather than to a bilayer; in qualitative terms the results can, however, be easily interpreted for the bilayer). Classical bilayer and multilayer structures that form in charged particle traps have also been studied theoretically by Dubin \[12a\] and have recently been observed in ion traps \[12b\]. There are two problematic issues that make the predictions based on the calculations less than reliable . The first issue relates to the approximation technique used: most of the works cited use methods which violate the $`3^{rd}`$ frequency-moment-sum-rule, whose satisfaction is well recognized to be an important criterion for providing an acceptable description of the collective mode behavior. The second issue concerns the use of the intralayer and interlayer correlation functions as inputs in all the calculations cited. No reliable pair correlation function (PCF) data - either for classical or quantum bilayer systems - have been available until fairly recently: thus predictions of the collective mode structure (which turns out to be extremely sensitive to the behavior of the inputted PCF) have been compromised from the outset.
In this Letter we present the first consistent and reliable calculation of the collective mode spectrum of a strongly coupled electronic bilayer liquid. Our results show features which are qualitatively different from the weakly coupled RPA results; they also show that earlier claims concerning the possible emergence of a dynamical instability cannot be supported by a more consistent treatment of the correlations. The calculation is based on a purely classical model: (i) two 2D electron liquids separated by distance $`d`$; (ii) scattering on impurities etc. neglected; (iii) no interlayer tunneling; (iv) the system is described as a binary liquid with interaction potentials $`\phi _{11}(r)=\phi _{22}(r)=e^2/r,\phi _{12}(r)=e^2/\sqrt{(r^2+d^2)}`$. In addition, the model portrays a classical electron liquid where exchange and other quantum effects are neglected: this approximation is reasonable in the strong coupling domain where the particles are well localized. The intralayer coupling is characterized by the parameter $`\mathrm{\Gamma }=e^2/(aT)`$ where $`a=1/\sqrt{(}\pi n)`$ and $`T`$ is the kinetic energy per particle โ the temperature in a classical system and $`(1/2)\epsilon _F`$ in a zero temperature 2D electron gas. Hence the equivalence $`\mathrm{\Gamma }2r_s`$. The calculation of the dielectric matrix $`\epsilon _{ij}^{\mu \nu }(๐ค\omega )`$ is carried out in the Quasi Localized Charge Approximation (QLCA), which has been applied successfully for the description of other strongly coupled Coulomb systems ; in the QLCA $`\epsilon (๐ค\omega )`$ becomes a functional of the intralayer and interlayer PCF-s $`h_{11}(r)`$ and $`h_{12}(r)`$ or of the corresponding structure functions $`S_{11}(k)`$ and $`S_{12}(k)`$:
$$\epsilon (๐ค\omega )=๐\omega _0^2(ka)[\omega ^2๐๐(๐ค)]^1$$
(1)
($`\omega _0=(\frac{2\pi e^2n}{ma})^{1/2}`$, the nominal plasma frequency of a single 2D layer) with
$`D_{ij}^{\mu \nu }(๐ค)=`$ (2)
$`{\displaystyle \frac{1}{mA}}{\displaystyle \underset{๐ช}{}}q^\mu q^\nu [\phi _{ij}(q)S_{ij}(|๐ค๐ช|)\delta _{ij}{\displaystyle \underset{l}{}}\phi _{il}(q)S_{il}(q)],`$ (3)
$`๐`$ is the identity matrix. In Cartesian space $`\epsilon (๐ค\omega )`$ and similarly $`๐(๐ค)`$ are reducible to longitudinal $`\epsilon ^L(๐^L)`$ and transverse $`\epsilon ^T(๐^T)`$ matrices. The dispersion relation for the longitudinal modes is then obtained from
$$\epsilon ^L(๐ค\omega )=0$$
(4)
which leads to
$$\omega ^2=\omega _0ka(1\pm e^{kd})+D_{11}^L(๐ค)\pm D_{12}^L(๐ค)$$
(5)
With the neglect of retardation effects, the dispersion relation for the transverse modes is derived from
$$\epsilon ^T(๐ค\omega )^1=0$$
(6)
which yields
$$\omega ^2=D_{11}^T(๐ค)\pm D_{12}^T(๐ค)$$
(7)
The $`๐`$-functions are to be expressed in terms of the structure functions, according to Eq. 3. These latter have recently been calculated through the HNC (Hypernetted Chain) integral equation for a wide range of $`\mathrm{\Gamma }`$ and $`d`$ values. Inputting them into the corresponding $`๐`$-functions and using the latter in Eqs. 5 and 7 one can generate a full description of the collective mode spectrum. The results are portrayed in Figs. 1-3 and the qualitative features of the collective mode dispersion are summarized below. (The figures are given for $`\mathrm{\Gamma }=40`$, corresponding to $`r_s=20`$: this $`r_s`$ value, while high enough for correlations to be dominant is within the domain of experimental realizability.)
1. The spectrum of collective modes comprises 4 modes: 2 (longitudinal and transverse) in-phase modes (corresponding to the + sign in Eqs. 5 and 7) and 2 (longitudinal and transverse) out-of-phase modes (corresponding to the - sign in Eqs. 5 and 7).
2. The in-phase modes (where the two layers oscillate in unison) are not qualitatively different from the similar modes of an isolated 2D layer . In particular, for $`k0`$ the longitudinal (plasmon) mode has the typical, quasi-acoustic $`\omega \sqrt{k}`$ dispersion, while the transverse (shear) mode is acoustic, $`\omega k`$; both modes are softened by intralayer and interlayer correlations.
3. The out-of-phase modes (where the oscillations of the two layers exhibit a $`180^{}`$ phase difference) are characterized by an $`\omega (k=0)>0`$ energy gap. The physical reason for the existence of an energy gap for layered systems has already been discussed elsewhere . The out-of-phase longitudinal mode in the RPA approximation has been identified as the acoustic plasmon since for $`k0`$, $`\omega 0`$ as $`k`$. The present calculation clearly shows the marked difference brought about by the strong correlations. Since at $`k=0`$ the isotropy of the system is unbroken, the plasmon and the shear modes share a common gap value.
4. From Eqs. 5 and 7 the gap value can be expressed as
$$\omega ^2(0,d)=\frac{\omega _0^2}{2}\underset{0}{\overset{\mathrm{}}{}}d(qa)(qa)^2e^{qd}S_{12}(q).$$
(8)
With increasing $`d`$ and consequently decreasing interlayer correlations, $`\omega (0)`$ shows a decreasing tendency and it virtually vanishes for $`d>1.5`$ when the separated layers become practically uncorrelated . This downward trend is, however, preceded by a slight upturn for $`0<d<0.16`$ (for $`\mathrm{\Gamma }30`$). Although the details of this behavior are not well understood, it is most likely due to the substitutional disorder that prevails in this region : the eigenfrequencies of the localized modes in the substitutionally disordered phase are expected to be higher than in the substitutionally ordered phase . Within the domain investigated, the $`\mathrm{\Gamma }`$-dependence of $`\omega (0)`$ is quite mild, but the QLCA being a strong coupling approximation, no inference concerning the behavior of $`\omega (0)`$ in the moderately coupled ($`\mathrm{\Gamma }<10`$) domain can be drawn from this observation. In fact, it is expected that for low enough $`\mathrm{\Gamma }`$ values, $`\omega (0)`$ tends to zero to match at $`\mathrm{\Gamma }=0`$ the predicted RPA behavior .
5. For finite $`k`$ values all the four dispersion curves develop an oscillatory behavior, generated by the similar behavior of the inputted structure functions. This behavior has also been identified for the isolated 2D layer \[16, 17b\]. The structure of the out-of-phase plasmon mode is of special interest here: the first sharp roton-like minimum has attracted attention in earlier studies which were based on the neglect or on a highly approximate treatment of the interlayer correlations. It was suggested that the minimum of $`\omega ^2`$ may dip below $`\omega ^2=0`$ \[5a, 9,11\] or may, at least, reach the close vicinity of $`\omega ^2=0`$ \[5b,5c\]. The former behavior would indicate a dynamical instability (heralding the onset of CDW-type ground state), \[5a, 11\]; the latter has been interpreted as the onset of a new high-$`k`$, low frequency mode \[5b, 5c\]. Our results show that the roton minimum never drops below the value already reached by the dispersion curve of the 2D layer. The consistent treatment of the interlayer and intralayer correlations thus precludes the existence of the effects conjectured in , virtually independently of any other approximation used.
6. At high k-values, for a given $`d`$ all the dispersion curves approach the same asymptotic frequency value, the frequency of a localized mode, a particle oscillating in the screening environment of the two layers:
$`\omega ^2(\mathrm{},d)`$ $`=`$ $`{\displaystyle \frac{1}{mA}}{\displaystyle \underset{๐ช}{}}[\phi _{11}(q)(S_{11}(q)1)+\phi _{12}(q)S_{12}(q)]q^2`$ (9)
$`=`$ $`{\displaystyle \frac{\omega _0^2}{4}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d(qa)(qa)^2[S_{11}(q)1+S_{12}(q)e^{qd}]`$ (10)
This result is, probably, only of academic interest, since it is unlikely that high-k modes would survive the damping mechanism operating in the system.
The QLCA is not geared to describe damping processes and therefore our calculation fails to provide information on the damping of the collective modes. However, some qualitative statements can be made. We concentrate on the out-of-phase modes only. There are three major damping mechanisms to be considered. These are (i) single pair excitation (Landau damping), (ii) multiple pair excitations and (iii) diffusive-migrational damping . The effect of the single pair excitations can be easily assessed from Figs. 1 and 2. Both of them show the pair excitation domain: it is clear that as long as the layer separation is not too large ($`d/a<1.5`$) for small k values both the out-of-phase plasmon and shear modes are well outside the continuum and are thus immune to Landau-damping. In a highly correlated plasma multi-particle excitations are, however, also operative, with increasing importance at higher $`k`$ values. The $`k\mathrm{}`$ portion of the dispersion curve emerging from the continuum would probably be heavily damped by this process. The diffusive-migrational shifting of the quasi-sites can originate from quasi-thermal diffusion or from tunneling between neighboring minima of the fluctuating potential. For $`\mathrm{\Gamma }`$ sufficiently high ($`\mathrm{\Gamma }40`$, i.e. $`r_s20`$, probably close to crystallization ), the latter effect should be significantly diminished. The former, however, is sizable when its characteristic time becomes comparable with the period of the mode under consideration. Based on our earlier estimate of the lowest surviving oscillation frequency, $`\omega _{min}`$ and on the MD results of Ref. (see also ) concerning the lowest propagating wave-number value for the shear mode, one can conclude that for $`\mathrm{\Gamma }=40`$ ($`r_s=20`$) the $`d/a<0.8`$ domain and for $`\mathrm{\Gamma }=20`$ ($`r_s=10`$) the $`d/a<0.6`$ domain can be safely assumend not to be seriously affected by this damping mechanism for either of the modes. (cf Fig. 1/a). For higher $`d`$ values the gap frequency $`\omega (0)`$ is below $`\omega _{min}`$ and thus the shear mode would be propagating for $`\omega >\omega _{min}`$ only; in the domain $`\omega <\omega _{min}`$ the longitudinal mode would be stripped of its correlational features and would revert to an RPA type acoustic plasmon with $`\omega (k0)0`$. (cf. Fig. 1/b).
The collective mode structure of the strongly coupled bilayer liquid presented in this Letter bears a close relationship to the phonon spectrum of the bilayer solid, recently calculated by Goldoni and Peeters . The four modes in the solid phase can be identified as the transverse acoustic and the longitudinal quasi-acoustic ($`\sqrt{k}`$) phonons and the transverse and longitudinal optical phonons. The โgapsโ exhibited by the latter are, in general, different because of the anisotropy of the lattice. In contrast, in the liquid state, there is only one single isotropic gap, as determined in this Letter. This liquid gap value is typically slightly above the arithmetic average of the optical frequencies of the transverse and longitudinal phonons, in the solid phase.
Concerning the possible observation of the features predicted in this Letter, one can suggest three main areas that should lend themselves to direct experimental verification: (1) the existence and the nonmonotonic $`d`$-dependence of the $`k=0`$ energy gap; (2) the existence of a transverse shear excitation with a high frequency and expected low damping (this is in a sharp contrast to the usual scenario for the shear mode in the liquid phase, which vanishes for $`k0`$ ; (3) the non-existence of the predicted instability or low frequency mode in the vicinity of the first roton minimum. We note that the reported Raman scattering experiments are inconclusive because of the low $`r_s`$ and relatively high $`k`$ values involved. Recent advances in fabricating high $`r_s`$ samples and small layer separation should render the suggested experiments feasible.
In summary, we have obtained a comprehensive picture of the collective mode structure of an electronic bilayer in the strongly coupled liquid phase. This structure is qualitatively different both from that of the weakly coupled bilayer electron gas (describable in the RPA) and from the phonon spectrum of the bilayer solid and exhibit a number of remarkable features which should be experimentally verifiable. The calculation avoids the pitfalls of earlier approaches which stem from the inconsistencies in the approximations used for the interlayer and for the intralayer PCF-s, and from the violation of the $`3^{rd}`$ frequency-moment-sum rule.
We wish to thank Dan Dubin for making his unpublished results available to us. GK is grateful to Tom OโNeil for making possible his stay at UCSD, where this work was started. We thank Hongbo Zhao for help in preparation of the Figures. This work has been partially supported by DOE Grants DE-FG02-98ER54501 and DE-FG02-98ER54491. |
warning/0003/quant-ph0003089.html | ar5iv | text | # Cavity induced modifications to the resonance fluorescence and probe absorption of a laser-dressed V atom
## I Introduction
A major interest of modern quantum optics is to devise ways to modify and control the radiative properties of atoms. This may be achieved by changing the environment, so that the atoms interact with a modified set of vacuum modes. One such modified vacuum is provided by the cavity environment , where the electromagnetic modes are concentrated around the cavity resonant frequency. The coupling of the atoms to the modified electromagnetic vacuum is therefore frequency dependent. For an excited atom located inside such a cavity, the cavity mode is the only one available to the atom for emission. If the atomic transition is in resonance with the cavity, the spontaneous emission rate into the particular cavity mode is enhanced ; otherwise, it is inhibited . When the atom is strongly driven by a laser field, the atom-laser system may be considered to form a new dressed atom whose energy-level structure is intensity dependent. For such a coherently driven two-level atom placed inside a cavity, theoretical investigations have predicted a phenomenological richness which is not found in the absence of the strong driving fieldโfor example, dynamical suppression of the spontaneous emission rate , population inversion in both bare and dressed states basis , and distortion and narrowing of the Mollow triplet . All these features are very sensitive to the cavity resonance frequency because of the cavity enhancement of the dressed atomic transitions.
Recently, Lange and Walther have observed the dynamical suppression of spontaneous emission in a microwave cavity. In the optical-frequency regime, Zhu et al. have also reported experimental studies of the effects of cavity detuning on the radiative properties of a coherently driven two-level atom. They have shown that the atomic fluorescence of a strongly driven two-level atom is enhanced when the cavity frequency is tuned to one of the sidebands of the Mollow fluorescence triplet, whereas it is inhibited by tuning to the other sideband. The enhancement of atomic resonance fluorescence at one sideband is a direct demonstration of population inversion.
In this paper, we investigate the dynamical modification of the resonance fluorescence of a coherently driven V-type three-level atom coupled to a frequency-tunable, single-mode cavity in the bad cavity limit. We demonstrate that the atomic populations, fluorescence spectrum and absorption spectrum can be strongly controlled and manipulated by tuning the cavity frequency. In Section 2, we derive a cavity-modified master equation for the atomic density-matrix operator from the full master equation by adiabatically eliminating the cavity variables in the bad cavity limit. For simplicity, we restrict attention to the situation where the laser frequency is tuned to the mean Bohr frequency of the excited states. The results obtained here are basic to the whole paper, and the equations derived in this section are used to calculate all the numerical results presented. The first part of Section 3 is devoted to discussing the atomic population distribution in the bare state representation. It is pointed out that significant population inversions can be achieved. In the second part of this section, we analyse the equations for the populations in the dressed state basis, with particular emphasis on the high field limit. We find that the dressed state populations obey rate equations, with atomic transition rates that are strongly dependent on the Rabi frequency and the cavity resonance frequency. Under certain conditions, the population may also be inverted in the dressed-state basis. The plots of the dressed state populations against the cavity frequencyโlaser frequency detuning are used to provide a semi-quantitative understanding of the phenomena described in the subsequent sections. In Section 4 we study the cavity effects on the resonance fluorescence spectra of this system. It is shown that dynamical control of the populations and fluorescence spectrum by adjustment of the cavity resonance frequency and Rabi frequency is possible. Peak suppression and line narrowing phenomena are also revealed. In Section 5 we briefly consider the absorption of a weak tunable probe field transmitted through this system, and in the last section we give a summary.
## II The cavity modified master equation
Consider a V-configuration atom consisting of two excited states $`|1`$ and $`|2`$ coupled to a ground state $`|0`$ by a single-mode cavity field of frequency $`\omega _C`$ and a laser field with frequency $`\omega _L`$, as shown in Fig. 1. The cavity mode is described by the annihilation and creation operators $`a`$ and $`a^{}`$, while the atom is represented by the operators $`A_{lk}|lk|(l,k=0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}2})`$. In the frame rotating at the frequency $`\omega _L,`$ and within the rotating wave approximation, the master equation for the density matrix operator $`\rho `$ of the combined atom-cavity system is
$$\dot{\rho }_T=i[H_A+H_C+H_I,\rho _T]+_A\rho _T+_C\rho _T$$
(1)
where
$`H_A=(\mathrm{\Delta }\omega _{21})A_{11}+\mathrm{\Delta }A_{22}+\mathrm{\Omega }_2(A_{02}+A_{20})+\mathrm{\Omega }_1(A_{01}+A_{10}),`$ (3)
$`H_C=\delta a^{}a,`$ (4)
$`H_I=g_2(a^{}A_{02}+A_{20}a)+g_1(a^{}A_{01}+A_{10}a),`$ (5)
$`_A\rho _T={\displaystyle \frac{\gamma _1}{2}}(2A_{01}\rho _TA_{10}\rho _TA_{11}A_{11}\rho _T)+{\displaystyle \frac{\gamma _2}{2}}(2A_{02}\rho _TA_{20}\rho _TA_{22}A_{22}\rho _T),`$ (6)
$`_C\rho _T=\kappa (2a\rho _Ta^{}a^{}a\rho _T\rho _Ta^{}a),`$ (7)
with
$$\omega _{21}=\omega _2\omega _1,\delta =\omega _C\omega _L\text{and}\mathrm{\Delta }=\omega _2\omega _L.$$
(8)
Here $`H_A`$ and $`H_C`$ describe the coherently driven atom and the cavity respectively, and $`H_I`$ represents the interaction between the atom and the cavity mode. The Rabi frequency $`\mathrm{\Omega }_j`$ relates to the atomic transitions $`|j|0`$ $`\left(j=1,2\right)`$ under the action of the driving laser field, and $`g_j`$ is the coupling constant between the atom and the cavity mode associated with the same transition. $`_C\rho _T`$ and $`_A\rho _T`$ describe respectively the damping of the cavity field by a standard vacuum reservoir, and the atomic damping to background modes other than the privileged cavity modes. $`\gamma _1`$ and $`\gamma _2`$ are just the spontaneous decay constants of the levels $`|1`$ and $`|2`$. Here we also assume that the atomic dipole moments $`๐_{10}`$ and $`๐_{20}`$ are orthogonal to each other, so that there is no spontaneously generated quantum interference resulting from the cross coupling between the transitions $`|1|0`$ and $`|2|0`$. For simplicity in the resulting expressions, we assume $`\mathrm{\Omega }_1=\mathrm{\Omega }_2=\mathrm{\Omega }`$, $`g_1=g_2=g`$, $`\gamma _1=\gamma _2=\gamma `$, and $`\mathrm{\Delta }=\omega _{21}/2`$ in what follows.
We assume that the atom-cavity coupling is weak and the cavity has a low-Q value, so that
$$\kappa g\gamma $$
(9)
(the bad cavity limit). This condition implies that the cavity mode response to the standard vacuum reservoir is much faster than that produced by its interaction with the atom. Then the atom always experiences the cavity mode in the state induced by the vacuum reservoir, and this permits one to eliminate the variables containing the cavity field operators adiabatically, giving rise to a reduced master equation for the atomic variables only. As the derivation is tedious, we refer readers to , and here only outline the key points.
We temporarily disregard $`_A\rho _T`$ in the elimination of the cavity-mode, since it unchanged by these operations. First we perform a canonical transformation to the atom-cavity interaction picture (1) by
$$\stackrel{~}{\rho }_T=e^{i(H_A+H_C)t}\rho _Te^{i(H_A+H_C)t}.$$
(10)
The master equation then takes the form
$$_t\left(e^{_Ct}\stackrel{~}{\rho }_T\right)=ie^{_Ct}[\stackrel{~}{H}_I(t),\stackrel{~}{\rho }_T],$$
(11)
where $`\stackrel{~}{H}_I(t)=g[\stackrel{~}{D}(t)a^{}\mathrm{exp}(i\delta t)+h.c.]`$, with $`\stackrel{~}{D}(t)=\mathrm{exp}(iH_At)D\mathrm{exp}(iH_At)`$ and $`D=A_{01}+A_{02}`$. We next introduce the operator $`\chi `$
$$\chi =e^{_Ct}\stackrel{~}{\rho }_T,$$
(12)
which, according to Eq. (11), obeys the equation
$`\dot{\chi }(t)`$ $`=`$ $`ige^{\kappa t}\left\{[a^{},\stackrel{~}{D}(t)\chi (t)]e^{i\delta t}+[a,\chi (t)\stackrel{~}{D}^{}(t)]e^{i\delta t}\right\}`$ (14)
$`ige^{\kappa t}\left\{[\stackrel{~}{D}(t),\chi (t)a^{}]e^{i\delta t}+[\stackrel{~}{D}^{}(t),a\chi (t)]e^{i\delta t}\right\}.`$
Due to the smallness of the coupling constant $`g`$, we can perform a second-order perturbation calculation with respect to $`g`$ by means of standard projection operator techniques. Noting that
$$\text{Tr}_C\chi (t)\text{Tr}_C\stackrel{~}{\rho }_T(t)\stackrel{~}{\rho }(t),$$
(15)
we can trace out the cavity variables to obtain the master equation for the reduced density matrix operator $`\stackrel{~}{\rho }`$ of the atom. Under the Born-Markovian approximation, the resulting master equation is of the form
$`\stackrel{~}{\rho }(t)=g^2{\displaystyle _0^{\mathrm{}}}\{e^{(\kappa +i\delta )\tau }[\stackrel{~}{D}^{}(t)\stackrel{~}{D}(t\tau )\stackrel{~}{\rho }(t)\stackrel{~}{D}(t\tau )\stackrel{~}{\rho }(t)\stackrel{~}{D}^{}(t)]+h.c.\}d\tau .`$
Finally transforming $`\stackrel{~}{\rho }`$ back to the original picture via $`\rho =\mathrm{exp}(iH_At)\stackrel{~}{\rho }\mathrm{exp}(iH_At)`$, and restoring the $`_A\rho `$ contribution, we express the reduced master equation for the atomic variables as
$`\dot{\rho }`$ $`=`$ $`i[H_A,\rho ]`$ (18)
$`+{\displaystyle \frac{\gamma _c}{2}}\left(D\rho S^{}+S\rho D^{}D^{}S\rho \rho S^{}D\right)`$
$`+{\displaystyle \frac{\gamma }{2}}\left(2A_{01}\rho A_{10}\rho A_{11}A_{11}\rho \right)+{\displaystyle \frac{\gamma }{2}}\left(2A_{02}\rho A_{20}\rho A_{22}A_{22}\rho \right)`$
where $`\gamma _c=2g^2/\kappa `$ specifies the emission rate of the atom into the cavity mode, and
$`S`$ $`=`$ $`\kappa {\displaystyle _0^{\mathrm{}}}e^{(\kappa +i\delta )\tau }\stackrel{~}{D}(\tau )๐\tau `$ (19)
$`=`$ $`\beta _0A_{00}+\beta _1A_{11}+\beta _2A_{22}+\beta _3A_{10}+\beta _4A_{01}+\beta _5A_{20}+\beta _6A_{02}+\beta _7A_{21}+\beta _8A_{12},`$ (20)
where the coefficients $`\beta _i`$ ($`i=0,1,\mathrm{}.8`$) are given by
$$\left[\begin{array}{c}\beta _0\hfill \\ \beta _1\hfill \\ \beta _2\hfill \\ \beta _3\hfill \\ \beta _4\hfill \\ \beta _5\hfill \\ \beta _6\hfill \\ \beta _7\hfill \\ \beta _8\hfill \end{array}\right]=\left[\begin{array}{ccccc}0& 2\eta \epsilon ^2& 2\eta \epsilon ^2& 8\eta ^3& 8\eta ^3\\ 2\eta \epsilon & \eta \epsilon \left(1\epsilon \right)& \eta \epsilon \left(1+\epsilon \right)& \eta \left(1\epsilon ^2\right)/2& \eta \left(1\epsilon ^2\right)/2\\ 2\eta \epsilon & \eta \epsilon \left(1+\epsilon \right)& \eta \epsilon \left(1\epsilon \right)& \eta \left(1\epsilon ^2\right)/2& \eta \left(1\epsilon ^2\right)/2\\ 4\eta ^2& 4\eta ^2\epsilon & 4\eta ^2\epsilon & 2\eta ^2\left(1+\epsilon \right)& 2\eta ^2\left(1\epsilon \right)\\ 4\eta ^2& \epsilon ^2\left(1\epsilon \right)/2& \epsilon ^2\left(1+\epsilon \right)/2& 2\eta ^2\left(1\epsilon \right)& 2\eta ^2\left(1+\epsilon \right)\\ 4\eta ^2& 4\eta ^2\epsilon & 4\eta ^2\epsilon & 2\eta ^2\left(1\epsilon \right)& 2\eta ^2\left(1+\epsilon \right)\\ 4\eta ^2& \epsilon \left(1+\epsilon \right)/2& \epsilon \left(1\epsilon \right)/2& 2\eta ^2\left(1+\epsilon \right)& 2\eta ^2\left(1\epsilon \right)\\ 0& \eta \epsilon \left(1\epsilon \right)& \eta \epsilon \left(1+\epsilon \right)& \eta \left(1\epsilon \right)^2/2& \eta \left(1+\epsilon \right)^2/2\\ 0& \eta \epsilon \left(1+\epsilon \right)& \eta \epsilon \left(1\epsilon \right)& \eta \left(1+\epsilon \right)^2/2& \eta \left(1\epsilon \right)^2/2\end{array}\right]\left[\begin{array}{c}\frac{\kappa }{\kappa +i\delta }\\ \\ \frac{\kappa }{\kappa +i\left(\delta \mathrm{\Omega }_R\right)}\\ \\ \frac{\kappa }{\kappa +i\left(\delta +\mathrm{\Omega }_R\right)}\\ \\ \frac{\kappa }{\kappa +i\left(\delta 2\mathrm{\Omega }_R\right)}\\ \\ \frac{\kappa }{\kappa +i\left(\delta +2\mathrm{\Omega }_R\right)}\end{array}\right],$$
(21)
with
$$\mathrm{\Omega }_R=\frac{1}{2}\sqrt{\omega _{21}^2+8\mathrm{\Omega }^2},\eta =\frac{\mathrm{\Omega }}{2\mathrm{\Omega }_R},\epsilon =\frac{\omega _{21}}{2\mathrm{\Omega }_R}.$$
(22)
The first term of eq. (18) describes the coherent evolution of the atom, the second terms the cavity-induced decay of the atom into the cavity mode, and the remaining terms the atomic spontaneous emissions to the background modes.
The equations of motion for the atomic variables take the form
$`\dot{\rho }_{11}`$ $`=`$ $`\gamma \rho _{11}i\mathrm{\Omega }\left(\rho _{01}\rho _{10}\right)`$ (24)
$`{\displaystyle \frac{\gamma _c}{2}}\left(\beta _0\rho _{01}+\beta _0^{}\rho _{10}+\beta _4\rho _{11}+\beta _4^{}\rho _{11}+\beta _6\rho _{21}+\beta _6^{}\rho _{12}\right),`$
$`\dot{\rho }_{22}`$ $`=`$ $`\gamma \rho _{22}i\mathrm{\Omega }\left(\rho _{02}\rho _{20}\right)`$ (26)
$`{\displaystyle \frac{\gamma _c}{2}}\left(\beta _0\rho _{02}+\beta _0^{}\rho _{20}+\beta _4\rho _{12}+\beta _4^{}\rho _{21}+\beta _6\rho _{22}+\beta _6^{}\rho _{22}\right),`$
$`\dot{\rho }_{10}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\gamma i\omega _{21}\right)\rho _{10}+i\mathrm{\Omega }\left(\rho _{11}\rho _{00}+\rho _{12}\right)`$ (28)
$`{\displaystyle \frac{\gamma _c}{2}}\left[\beta _0\rho _{00}\beta _1\left(\rho _{11}+\rho _{12}\right)\beta _3\left(\rho _{01}+\rho _{02}\right)+\beta _4\rho _{10}+\beta _6\rho _{20}\beta _8\left(\rho _{21}+\rho _{22}\right)\right],`$
$`\dot{\rho }_{20}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\gamma +i\omega _{21}\right)\rho _{20}+i\mathrm{\Omega }\left(\rho _{22}\rho _{00}+\rho _{21}\right)`$ (30)
$`{\displaystyle \frac{\gamma _c}{2}}\left[\beta _0\rho _{00}\beta _2\left(\rho _{21}+\rho _{22}\right)\beta _5\left(\rho _{01}+\rho _{02}\right)+\beta _4\rho _{10}+\beta _6\rho _{20}\beta _7\left(\rho _{11}+\rho _{12}\right)\right],`$
$`\dot{\rho }_{21}`$ $`=`$ $`\left(\gamma +i\omega _{21}\right)\rho _{21}i\mathrm{\Omega }\left(\rho _{01}\rho _{20}\right)`$ (32)
$`{\displaystyle \frac{\gamma _c}{2}}\left(\beta _0\rho _{01}+\beta _0^{}\rho _{20}+\beta _4\rho _{11}+\beta _4^{}\rho _{21}+\beta _6\rho _{21}+\beta _6^{}\rho _{22}\right),`$
## III Steady-state populations
### A The bare states.
We have numerically solved the equations (32) for the populations in the steady state. In all our numerical plots we assume the values $`\gamma =1,g=20`$ and $`\kappa =100,`$ so that the condition (9) is satisfied. All the frequencies are also measured in units of $`\gamma .`$
In Fig. 2 we plot the bare state populations as a function of the cavity-laser detuning $`\delta .`$ We first consider the case where $`\omega _{21}=10,`$ taking in frame (a) $`\mathrm{\Omega }=4;`$ in frame (b), $`\mathrm{\Omega }=10;`$ and in frame (c), $`\mathrm{\Omega }=100`$. For the smallest value of $`\mathrm{\Omega },`$ all three populations tend to roughly the same value for large detunings, but for small detunings, a resonant effect is evident around $`\delta =0:`$ the population in the ground state passes through a maximum, whilst the population in the two excited states exhibits a minimum. The behaviour in frame (b) is qualitatively different: here, the population resides mainly in the ground state, but this population now shows a dip as $`\delta `$ passes through zero. The excited state populations show a minimum and a maximum close to the origin. There is a very tiny amount of population inversion for state $`|2`$ over state $`|0`$ for a very short range of negative detunings close to zero, but the effect is unimportant. As the value of $`\mathrm{\Omega }`$ is increased to $`\mathrm{\Omega }=100`$ in frame (c), keeping $`\omega _{21}=10,`$ the populations show a flatter behaviour, with the ground state population tending to the value 0.5 for large $`\delta ,`$ and that of the excited states to the value 0.25. The minimum in the ground state population at $`\delta =0`$ in frame (b) has been replaced by a very shallow maximum in frame (c). It is clear that the most interesting behaviour arises for $`\mathrm{\Omega }\omega _{21}.`$
In the next three frames we assume a larger value for the excited state splitting, $`\omega _{21}=200.`$ For the lowest value of $`\mathrm{\Omega }`$ considered, $`\mathrm{\Omega }=100,`$ the behaviour in frame (d) is qualitatively different to that shown in the first three frames. Now all three bare populations tend towards the same value for large $`\delta .`$ The ground state population is almost flat, but the excited states possess pronounced maxima and minima, as well as the suggestion of further structure. It is clear that, for appropriate detunings, a large population inversion between the excited and ground states can be achieved. When $`\mathrm{\Omega }`$ is increased to $`\mathrm{\Omega }=200,`$ as in frame (e), the ground state population begins to increase, decreasing the population inversion obtainable. Also structure at four different frequencies becomes evident. (The reasons for this will become apparent in the following subsection.) The trend continues in the final frame, where $`\mathrm{\Omega }=300,`$ and there is no population inversion. For still larger values of $`\mathrm{\Omega }`$ (not shown), the ground state population tends to flatten around the value 0.5, and the excited state populations around the value 0.25, as in frame (c) but with different detailed structure.
### B The dressed states.
To study the modification of the atomic populations due to the presence of the cavity, we work in the semiclassical dressed-state representation. The dressed states, defined by the eigenvalue equation, $`H_A|\alpha =\lambda _\alpha |\alpha `$, are of the form
$`|a={\displaystyle \frac{1}{2}}[(1\epsilon )|2(1+\epsilon )|1+4\eta |0],`$ (33)
$`|b=2\eta |2+2\eta |1+\epsilon |0,`$ (34)
$`|c={\displaystyle \frac{1}{2}}[(1+\epsilon )|2+(1\epsilon )|1+4\eta |0],`$ (35)
and the corresponding energies are
$$\lambda _a=\mathrm{\Omega }_R,\lambda _b=0,\lambda _c=\mathrm{\Omega }_R.$$
(36)
Within the secular approximation, the equations of motion for the populations in the dressed states may be cast into the rate equation form
$`\dot{\rho }_{aa}`$ $`=`$ $`(R_{ab}+R_{ac})\rho _{aa}+R_{ca}\rho _{cc}+R_{ba}\rho _{bb}`$ (37)
$`\dot{\rho }_{bb}`$ $`=`$ $`(R_{bc}+R_{ba})\rho _{bb}+R_{cb}\rho _{cc}+R_{ab}\rho _{aa}`$ (38)
$`\dot{\rho }_{cc}`$ $`=`$ $`(R_{ca}+R_{cb})\rho _{cc}+R_{ac}\rho _{aa}+R_{bc}\rho _{bb}`$ (39)
where $`R_{\alpha \beta }(\alpha ,\beta =a,b,c)`$ represents the atomic transition rate from the substate $`|\alpha `$ of one dressed-state triplet to the substate $`|\beta `$ of the dressed-state triplet below, as depicted in Fig. 3. In the high field limit, that is, when the effective Rabi frequency is much greater than all the relaxation rates, $`\mathrm{\Omega }_R\gamma ,\gamma _c`$, the coupling between atomic density matrix elements $`\rho _{\alpha \beta }`$ associated with the various frequencies may be omitted to $`O(\gamma /\mathrm{\Omega }_R)`$ and $`O(\gamma _c/\mathrm{\Omega }_R)`$, and the transition rates may be expressed as
$`R_{bc}=R_{ba}={\displaystyle \frac{\gamma }{2}}(1\epsilon ^2)^2,`$ (40)
$`R_{ab}={\displaystyle \frac{\gamma }{2}}(1+\epsilon ^2)\epsilon ^2+\gamma _c\epsilon ^2(\mathrm{\Omega }_R),`$ (41)
$`R_{cb}={\displaystyle \frac{\gamma }{2}}(1+\epsilon ^2)\epsilon ^2+\gamma _c\epsilon ^2(\mathrm{\Omega }_R),`$ (42)
$`R_{ac}={\displaystyle \frac{\gamma }{4}}(1\epsilon ^4)+\gamma _c4\eta ^2(2\mathrm{\Omega }_R),`$ (43)
$`R_{ca}={\displaystyle \frac{\gamma }{4}}(1\epsilon ^4)+\gamma _c4\eta ^2(2\mathrm{\Omega }_R),`$ (44)
where $`(\pm x)=\kappa ^2/\left[\kappa ^2+(\delta \pm x)^2\right].`$ The above expressions for the transition rates will provide the basis for physical explanation of the effects to be described.
The function $`(\pm 2\mathrm{\Omega }_R)`$ introduces resonances at $`\delta =\pm 2\mathrm{\Omega }_R`$ into the rates $`R_{ac}`$ and $`R_{ca}`$ with a strength proportional to $`4\mathrm{\Omega }^2.`$ This appears to be more dominant than the resonances at $`\delta =\pm \mathrm{\Omega }_R`$ which occur in $`R_{cb}`$ and $`R_{ab}`$ with a strength proportional to $`\omega _{21}^2.`$ Resonances at $`\delta =\pm 2\mathrm{\Omega }_R`$ can be clearly seen in frames (b) to (f) of Fig. 2, whereas resonances at $`\delta =\pm \mathrm{\Omega }_R`$ are only apparent in frames (c) โ (f).
The equations (44) show that in the presence of the cavity, the transition rates $`R_{ab}`$, $`R_{ac}`$, $`R_{cb}`$ and $`R_{ca}`$ are strongly dependent on the cavity frequency, but $`R_{bc}`$ and $`R_{ba}`$ are only related to the spontaneous emission rate. This is because for the system considered here, a V-type three-level atom interacting with the cavity mode, there exist nine double-channels for the atomic transition from the dressed states $`|i`$ to $`|j`$, which originate from the atomic bare state transitions $`|1`$ to $`|0`$ and $`|2`$ to $`|0`$ respectively. Constructive or destructive interference happens within every double-channel. But the two double-channels for the transitions $`|b|c`$ and $`|b|a`$ are completely destructive for the cavity modeโthat is, the transitions from $`|b`$ to $`|c`$ and from $`|b`$ to $`|a`$ never result in the emission of a photon into the cavity mode. So the atomic transition rates $`R_{bc}`$ and $`R_{ba}`$ are only dependent on the spontaneous emission rate and independent of the cavity frequency. As the other seven double-channels are still open to the cavity mode, this means that for the other transitions from $`|i`$ to $`|j`$, a cavity photon can be generated, but it will be very strongly damped under the bad-cavity assumption. Therefore, besides the terms describing atomic photon emission into the background, there occur additional terms dependent on the cavity frequency in the transition rates $`R_{ab}`$, $`R_{ac}`$, $`R_{cb}`$ and $`R_{ca}`$.
The steady-state dressed populations are found from equations (39) to be
$`\rho _{aa}={\displaystyle \frac{R_{ba}(R_{ca}+R_{cb}+R_{bc})+R_{bc}(R_{ca}R_{ba})}{(R_{ab}+R_{ac}+R_{ba})(R_{ca}+R_{cb}+R_{bc})(R_{ca}R_{ba})(R_{ac}R_{bc})}},`$ (45)
$`\rho _{cc}={\displaystyle \frac{R_{ba}(R_{ac}R_{bc})+R_{bc}(R_{ab}+R_{ac}+R_{ba})}{(R_{ab}+R_{ac}+R_{ba})(R_{ca}+R_{cb}+R_{bc})(R_{ca}R_{ba})(R_{ac}R_{bc})}},`$ (46)
$`\rho _{bb}=1\rho _{aa}\rho _{cc}.`$ (47)
The dressed state populations are plotted as a function of the detuning in Fig. 4, for the same parameter values employed in Fig. 3. These plots have been obtained by a numerical solution of equations (32), but we have found that the equations (47) provide an excellent approximation in the strong field limit.
It is evident that if the excited levels of the V-atom are degenerate ($`\epsilon =0`$) or nearly degenerate ($`\epsilon 0`$), the transitions $`|a|b`$ and $`|c|b`$ are turned off (as $`R_{ab},R_{cb}\epsilon ^2`$), whereas the rate of transitions out of the dressed state $`|b`$, $`R_{ba}+R_{bc},`$ is nonzero. There is thus no steady-state population in the dressed state $`|b`$, only in the dressed states $`|a`$ and $`|c`$. In general, it follows that in the regime $`\mathrm{\Omega }^2\omega _{21}^2,`$ the population in dressed state $`|b`$ is very small. Furthermore, if the cavity frequency is tuned to $`\delta =2\mathrm{\Omega }_R`$, the dressed state transition from $`|a`$ to $`|c`$ is resonantly enhanced ($`R_{ac}\gamma /4+\gamma _c4\eta ^2`$), while the reverse transition is suppressed, $`R_{ca}\gamma /4`$, and thus more population will be accumulated into the state $`|c`$. For $`\delta =2\mathrm{\Omega }_R`$, there exists a greater population in the dressed state $`|a`$. An example is shown in frame (c) of Fig. 4, where we take $`\omega _{21}=10,\mathrm{\Omega }=100.`$ (The population in $`|b`$ is very close to zero.) In frame (f), where we are only beginning to approach the limit $`\mathrm{\Omega }^2\omega _{21}^2,`$ the behaviour is similar, but now there is a small but significant population in state $`|b.`$
However, in the opposite limit, where the excited level splitting is much greater than the Rabi frequency ($`\omega _{21}^2\mathrm{\Omega }^2`$) the transition rates $`R_{ac},R_{ca},R_{ba}`$ and $`R_{bc}`$ are very small, and the transitions into the dressed state $`|b`$, represented by $`R_{ab}`$ and $`R_{cb}`$, dominate. Eventually, the population in the dressed state $`|b`$ approaches unity, and that in the states $`|a`$ and $`|c`$ is very small. In this case, the resonance features of the dressed state populations regarding the cavity frequency are less pronounced. We do not plot this case here, but the features can be seen beginning to emerge in frame (a), and to a lesser extent, in frame (c). The intermediate regime, where $`\omega _{21}=\mathrm{\Omega },`$ is shown in frames (b) and (e).
In general, the population distributions are strongly dependent on the cavity frequency. For example, when $`\delta =0,`$ the cavity is tuned to resonance with the driving field, we have $`R_{ab}=R_{cb}`$, $`R_{ba}=R_{bc}`$ and $`R_{ac}=R_{ca}`$. Consequently, there is no population difference between the upper dressed state $`|c`$ and the lower one $`|a`$. Moreover, in the case $`\mathrm{\Omega }_R\kappa `$, the distribution is same as in free space .
When the cavity frequency satisfies $`\delta =\mathrm{\Omega }_R`$, describing resonance with the dressed state transition $`|a|b`$, the rate of this transition $`R_{ab}`$ is greatly enhanced. Also the rate of downward atomic transitions from $`|a`$ to $`|c`$ is larger than the transition rate $`|c|a`$, i.e., $`R_{ac}>R_{ca}`$. As a result, the population in the dressed state $`|c`$ is greater than that in the state $`|a,`$ $`(\rho _{cc}>\rho _{aa}).`$
A similar analysis shows that the population, $`\rho _{cc}`$, of the state $`|c`$ is also greater than the population, $`\rho _{aa}`$, of the dressed state $`|a`$ if the cavity is tuned to $`\delta =2\mathrm{\Omega }_R`$. The opposite conclusions hold when $`\delta =\mathrm{\Omega }_R`$ and $`2\mathrm{\Omega }_R`$, as is clearly demonstrated in Figs. 4(c) โ 4(f) for different values of the Rabi frequency $`\mathrm{\Omega }`$.
If the cavity is tuned to resonance with the driving field $`(\delta =0),`$ then $`R_{ab}=R_{cb},`$ $`R_{ca}=R_{ac}`$ and $`R_{bc}=R_{ba}`$. From Fig. 2 we see that the population distribution between the dressed states $`|c`$ and $`|a`$ is balanced, which is similar to the case in the absence of cavity . In the presence of the cavity, the transition rates $`R_{ab}`$ and $`R_{cb}`$from the dressed states $`|a`$ and $`|c`$ to the dressed state $`|b`$ are faster than those in the absence of the cavity, whilst the reverse rates $`R_{bc}`$ and $`R_{ba}`$ remain unchanged because of the closing of the atomic transitions $`|b|c`$ and $`|b|a`$ for the cavity mode due to destructive interference. The populations $`\rho _{cc}`$ and $`\rho _{aa}`$ are decreased and $`\rho _{bb}`$ is increased by comparison with the case in the absence of the cavity.
However, when the cavity frequency is tuned to satisfy $`\delta =\mathrm{\Omega }_R`$, the atomic transition from $`|a`$ to $`|b`$, which occurs at frequency $`\omega _L\mathrm{\Omega }_R,`$ is resonant with the cavity. The interaction of the atom with the privileged cavity mode is enhanced, and the atom predominately (since $`g\gamma `$) emits a photon into the cavity mode, characterized by the $`\epsilon ^2\gamma _c`$ term in $`R_{ab}`$, besides radiating a photon into the background, represented by the term $`\mathrm{\Gamma }_0`$ in $`R_{ab}`$. By contrast, the other transitions describing atomic emission of a photon into the cavity mode are far off resonance with the cavity (since $`\mathrm{\Omega }_R\kappa `$), so the atom can only radiate a photon into the background. As a consequence, the symmetry of the transitions from $`|c`$ to $`|b`$and from $`|a`$ to $`|b`$ is broken, which results in an enhanced population in the dressed state $`|c`$ over that in $`|a`$.
When the cavity frequency is tuned to satisfy $`\delta =2\mathrm{\Omega }_R`$, the atomic transition rate $`R_{ac}`$ at the cavity frequency $`\omega _L2\mathrm{\Omega }_R`$ is resonantly enhanced, and the other three rates $`R_{ab}`$, $`R_{cb}`$ and $`R_{ca}`$ are decreased. In this case, if $`\gamma _c\gamma \left(1+\epsilon ^2\right)/\left(1\epsilon ^2\right)`$, the other five transition rates are much smaller than $`R_{ca}`$. Thus any population in the dressed state $`|a,`$ resulting from transitions from $`|b`$ and $`|c`$ to $`|a,`$ will return to $`|c`$ very quickly because $`R_{ac}`$ is much greater than the other transition rates. So it seems that the atom is trapped in the two dressed states $`|b`$ and $`|c`$, and the population in the state $`|a`$ approaches zero. Similar explanations can be adopted for the cases $`\delta =\mathrm{\Omega }_R`$ and $`\delta =2\mathrm{\Omega }_R`$.
## IV Resonance fluorescence spectrum
The spectrum of the atomic fluorescence emission emitted from the side of the cavity is proportional to the Fourier transform of the steady-state correlation function $`lim_t\mathrm{}\stackrel{}{E}^{()}(\stackrel{}{r},t+\tau )\stackrel{}{E}^{(+)}(\stackrel{}{r},t)`$, where $`\stackrel{}{E}^{(\pm )}(\stackrel{}{r},t)`$ are the positive and negative frequency parts of the radiation field in the far zone, which consists of a free-field operator and a source field that is proportional to the atomic polarization operator. The fluorescence spectrum is given by
$$\mathrm{\Lambda }(\omega )=Re\underset{0}{\overset{\mathrm{}}{}}[A_{20}(t+\tau ),A_{02}(t)+A_{10}(t+\tau ),A_{01}(t)]_t\mathrm{}e^{i\omega \tau }๐\tau .$$
(48)
Because we have assumed that the atomic dipole moment elements $`\stackrel{}{d}_{20}`$ and $`\stackrel{}{d}_{10}`$ are perpendicular to each other, the two correlation functions $`A_{10}(t+\tau ),A_{02}(t)`$ and $`A_{20}(t+\tau ),A_{01}(t)`$ make no contribution to the fluorescence emission spectrum. Using the quantum regression theorem, the spectrum may be expressed as
$$\mathrm{\Lambda }(\omega )=Re\left[\stackrel{~}{F}_{01}\left(z\right)+\stackrel{~}{G}_{02}\left(z\right)\left(\left|\rho _{01}\left(\mathrm{}\right)\right|^2+\left|\rho _{02}\left(\mathrm{}\right)\right|^2\right)/z\right]_{z=i\omega }$$
(49)
where $`\stackrel{~}{X}\left(z\right)`$ denotes the Laplace transform of $`X\left(\tau \right)`$, and
$$F\left(\tau \right)=U^{}\left(\tau \right)\left[|01|\rho \left(\mathrm{}\right)\right]U\left(\tau \right)\text{and }G\left(\tau \right)=U^{}\left(\tau \right)\left[|02|\rho \left(\mathrm{}\right)\right]U\left(\tau \right)$$
(50)
with $`U\left(t\right)`$ the time development operator, obey the same equation of motion as $`\rho \left(\tau \right)`$ but with the different initial conditions implied by the definitions (50).
We calculate the resonance fluorescence spectra numerically from the equations obtained in Section II. However, the dressed atom approach provides a convenient way of interpreting the results so obtained, at least in the strong field limit, and so we develop this approach here. In terms of the dressed states (35), we have, for example,
$$\stackrel{~}{F}_{01}\left(z\right)=\underset{\alpha ,\beta =a,b,c}{}0|\alpha \stackrel{~}{F}_{\alpha \beta }\left(z\right)\beta |1$$
(51)
with the initial condition
$$F_{\alpha \beta }\left(0\right)=\underset{\gamma }{}\alpha |0\rho _{\gamma \beta }\left(\mathrm{}\right)1|\beta .$$
(52)
In the high field limit, $`\mathrm{\Omega }_R\gamma ,\gamma _c`$, which is the condition for the secular approximation to hold, the dressed energy levels are well separated, and one can associate the diagonal terms on the right hand side of equation (51), and those in the corresponding equation for $`\stackrel{~}{G}_{02}\left(z\right),`$ with the resonance fluorescence line at the centre of the spectrum,$`\omega =\omega _L`$ (see Fig. 3). In a similar way, the elements $`\stackrel{~}{F}_{ab},\stackrel{~}{G}_{ab},\stackrel{~}{F}_{bc}`$ and $`\stackrel{~}{G}_{bc}`$ are associated with the line at $`\omega =\omega _L\mathrm{\Omega }_R`$, whilst the elements $`\stackrel{~}{F}_{ba},\stackrel{~}{G}_{ba},\stackrel{~}{F}_{cbc}`$ and $`\stackrel{~}{G}_{cb}`$ are associated with the line at $`\omega =\omega _L+\mathrm{\Omega }_R.`$ Finally, $`\stackrel{~}{F}_{ac}`$ and $`\stackrel{~}{G}_{ac}`$ are associated with the line at $`\omega =\omega _L+2\mathrm{\Omega }_R,`$ and $`\stackrel{~}{F}_{ca}`$ and $`\stackrel{~}{G}_{ca}`$ with the line at $`\omega =\omega _L2\mathrm{\Omega }_R`$
One can apply the secular approximation to simplify the equations of motion for the atomic density matrix elements in the dressed state representation, which are obtained according to eq. (18) as follows
$`\dot{\rho }_{aa}=\mathrm{\Gamma }_{1a}\rho _{aa}+\mathrm{\Gamma }_{2a}\rho _{cc}+R_{ba},`$ (53)
$`\dot{\rho }_{cc}=\mathrm{\Gamma }_{1b}\rho _{cc}+\mathrm{\Gamma }_{2b}\rho _{aa}+R_{bc},`$ (54)
$`\dot{\rho }_{ab}=\left(\mathrm{\Gamma }_{3a}i\mathrm{\Omega }_3\right)\rho _{ab}+\mathrm{\Gamma }_4\rho _{bc},`$ (55)
$`\dot{\rho }_{bc}=\left(\mathrm{\Gamma }_{3b}i\mathrm{\Omega }_4\right)\rho _{bc}+\mathrm{\Gamma }_4\rho _{ab},`$ (56)
$`\dot{\rho }_{ac}=\left(\mathrm{\Gamma }_5i\mathrm{\Omega }_5\right)\rho _{ac},`$ (57)
with
$`\mathrm{\Gamma }_{1a}={\displaystyle \frac{\gamma }{4}}(32\epsilon ^2+3\epsilon ^4)+\gamma _c\left[\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (58)
$`\mathrm{\Gamma }_{1b}={\displaystyle \frac{\gamma }{4}}(32\epsilon ^2+3\epsilon ^4)+\gamma _c\left[\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (59)
$`\mathrm{\Gamma }_{2a}={\displaystyle \frac{\gamma }{2}}\left(1\epsilon ^2\right)\left(3\epsilon ^21\right)+\gamma _c4\eta ^2(2\mathrm{\Omega }_R),`$ (60)
$`\mathrm{\Gamma }_{2b}={\displaystyle \frac{\gamma }{2}}\left(1\epsilon ^2\right)\left(3\epsilon ^21\right)+\gamma _c4\eta ^2(2\mathrm{\Omega }_R),`$ (61)
$`\mathrm{\Gamma }_{3a}={\displaystyle \frac{\gamma }{4}}(3+\epsilon ^22\epsilon ^4)+{\displaystyle \frac{\gamma _c}{2}}\left[4\eta ^2(0)+\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (62)
$`\mathrm{\Gamma }_{3b}={\displaystyle \frac{\gamma }{4}}(3+\epsilon ^22\epsilon ^4)+{\displaystyle \frac{\gamma _c}{2}}\left[4\eta ^2(0)+\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (63)
$`\mathrm{\Gamma }_4={\displaystyle \frac{\gamma }{2}}\epsilon ^2(1\epsilon ^2),`$ (64)
$`\mathrm{\Gamma }_5={\displaystyle \frac{\gamma }{4}}(3+\epsilon ^4)+{\displaystyle \frac{\gamma _c}{2}}\left\{16\eta ^2(0)+\epsilon ^2\left[(\mathrm{\Omega }_R)+(\mathrm{\Omega }_R)\right]+4\eta ^2\left[(2\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right]\right\},`$ (65)
$`\mathrm{\Omega }_3=\mathrm{\Omega }_R+{\displaystyle \frac{\gamma _c}{2}}\left[4\eta ^2(0)+\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (66)
$`\mathrm{\Omega }_4=\mathrm{\Omega }_R+{\displaystyle \frac{\gamma _c}{2}}\left[4\eta ^2(0)+\epsilon ^2(\mathrm{\Omega }_R)+4\eta ^2(2\mathrm{\Omega }_R)\right],`$ (67)
$`\mathrm{\Omega }_5=2\mathrm{\Omega }_R+{\displaystyle \frac{\gamma _c}{2}}\{\epsilon ^2\left[(\mathrm{\Omega }_R)(\mathrm{\Omega }_R)\right]+4\eta ^2\left[(2\mathrm{\Omega }_R)(2\mathrm{\Omega }_R)\right]\},`$ (68)
where $`(\pm x)=\kappa (\delta \pm x)/\left[\kappa ^2+(\delta \pm x)^2\right]`$, $`\mathrm{\Gamma }_i`$ is the decay rate in the dressed state representation, which is dependent on the cavity frequency, while $`\mathrm{\Omega }_R\mathrm{\Omega }_{3,4}`$ and $`2\mathrm{\Omega }_R\mathrm{\Omega }_5`$ are the cavity-induced level shifts. In the bad cavity and high field limits, the shifts are negligibly small.
In the dressed state representation, the underlying physical processes are very transparent. As argued in the paragraph following equation (51), the downward transitions between the same dressed states of two adjacent dressed-state triplets give rise to the central component of the fluorescence spectrum, i.e.,
$$\mathrm{\Lambda }_0(\omega )=Re\left[\frac{N_0(z)}{\left(z+\mathrm{\Gamma }_{1a}\right)\left(z+\mathrm{\Gamma }_{1b}\right)\mathrm{\Gamma }_{2a}\mathrm{\Gamma }_{2b}}\right]_{z=i\omega },$$
(69)
with
$`N_0(z)`$ $`=`$ $`4\eta ^2\left(2z+\mathrm{\Gamma }_{1a}+\mathrm{\Gamma }_{1b}\mathrm{\Gamma }_{2a}\mathrm{\Gamma }_{2b}\right)\rho _{aa}\rho _{cc}2\eta ^2\left(19\epsilon ^2\right)\left[\mathrm{\Gamma }_{2a}\rho _{cc}+\mathrm{\Gamma }_{2b}\rho _{aa}\right]\rho _{bb}`$ (71)
$`+2\eta ^2\left(1+9\epsilon ^2\right)\left[\left(z+\mathrm{\Gamma }_{1a}\right)\rho _{cc}+\left(z+\mathrm{\Gamma }_{1b}\right)\rho _{aa}\right]\rho _{bb}.`$
This spectral component consists of two Lorentzians with linewidths $`2\gamma _0^\pm =\left(\mathrm{\Gamma }_{1a}+\mathrm{\Gamma }_{1b}\right)\pm \sqrt{\left(\mathrm{\Gamma }_{1a}\mathrm{\Gamma }_{1b}\right)^2+4\mathrm{\Gamma }_{2a}\mathrm{\Gamma }_{2b}}`$.
However, the downward transitions $`|a|b`$and $`|b|c`$ from one dressed-state triplet to the next triplet lead to the lower-frequency inner sideband, yielding an expression of the form
$$\mathrm{\Lambda }_1(\omega )=Re\left[\frac{4\eta ^2\left[8\eta ^2\left(z+\mathrm{\Gamma }_{3a}+i\mathrm{\Omega }_3\right)\epsilon ^2\mathrm{\Gamma }_4\right]\rho _{bb}+\frac{1}{2}\epsilon ^2\left[\left(1+\epsilon ^2\right)\left(z+\mathrm{\Gamma }_{3b}+i\mathrm{\Omega }_4\right)8\eta ^2\mathrm{\Gamma }_4\right]\rho _{aa}}{\left(z+\mathrm{\Gamma }_{3a}+i\mathrm{\Omega }_3\right)\left(z+\mathrm{\Gamma }_{3b}+i\mathrm{\Omega }_4\right)\mathrm{\Gamma }_4^2}\right]_{z=i\omega },$$
(72)
whilst the transitions $`|b|a`$ and $`|c|b`$ between two near-lying dressed-state triplets result in the higher-frequency inner sideband,
$$\mathrm{\Lambda }_2(\omega )=Re\left[\frac{4\eta ^2\left[8\eta ^2\left(z+\mathrm{\Gamma }_{3b}i\mathrm{\Omega }_4\right)\epsilon ^2\mathrm{\Gamma }_4\right]\rho _{bb}+\frac{1}{2}\epsilon ^2\left[\left(1+\epsilon ^2\right)\left(z+\mathrm{\Gamma }_{3a}i\mathrm{\Omega }_3\right)8\eta ^2\mathrm{\Gamma }_4\right]\rho _{cc}}{\left(z+\mathrm{\Gamma }_{3a}i\mathrm{\Omega }_3\right)\left(z+\mathrm{\Gamma }_{3b}i\mathrm{\Omega }_4\right)\mathrm{\Gamma }_4^2}\right]_{z=i\omega }.$$
(73)
Since the cavity-induced level shifts are negligible, $`\mathrm{\Lambda }_1`$ will display a single spectral line located at frequency $`\omega _L\mathrm{\Omega }_R`$, and $`\mathrm{\Lambda }_2`$ a line at $`\omega _L+\mathrm{\Omega }_R`$. It is evident that the inner sidebands are also composed of two Lorentzians with linewidths $`2\gamma _1^\pm =\left(\mathrm{\Gamma }_{3a}+\mathrm{\Gamma }_{3b}\right)\pm \sqrt{\left(\mathrm{\Gamma }_{3a}\mathrm{\Gamma }_{3b}\right)^2+4\mathrm{\Gamma }_4^2}`$.
The final transitions, $`|a|c`$ and $`|a|c,`$ respectively generate the lower-frequency and higher-frequency spectral lines of the outer sidebands, which are given by
$`\mathrm{\Lambda }_3(\omega )`$ $`=`$ $`Re\left[{\displaystyle \frac{2\eta ^2\left(1+\epsilon ^2\right)\rho _{aa}}{\left(z+\mathrm{\Gamma }_5+i\mathrm{\Omega }_5\right)}}\right]_{z=i\omega },`$ (74)
$`\mathrm{\Lambda }_4(\omega )`$ $`=`$ $`Re\left[{\displaystyle \frac{2\eta ^2\left(1+\epsilon ^2\right)\rho _{cc}}{\left(z+\mathrm{\Gamma }_5i\mathrm{\Omega }_5\right)}}\right]_{z=i\omega }.`$ (75)
The spectral lines are centred at frequencies $`\omega _L\pm 2\mathrm{\Omega }_R`$, respectively, and have width $`2\mathrm{\Gamma }_5`$. In equations (69) โ (75), the $`\rho _{jj}`$ which appear are the steady state dressed state occupation probabilities, given by equations (47)
We first consider the fluorescence spectrum of the atom with two degenerate (or near-degenerate) excited states, where the population of the dressed state $`|b`$ is negligible, as illustrated in Fig. 5, where $`\omega _{21}=10`$ and $`\mathrm{\Omega }=100,`$ and we consider different detunings. The results may be understood from equations (69) โ (75). It is not difficult to see from eqs. (72) and (73) that the inner sidebands $`\mathrm{\Lambda }_{1,2}`$ will be very small, as $`\rho _{bb}0`$ and $`\epsilon ^20`$. Hence the central line and the outer sidebands will dominate. This is evident in all the spectra of Fig. 5, but particularly for frame (a) where the inner sidebands are hardly visible. When the cavity frequency $`\delta =0`$, the spectrum is a symmetric, Mollow-like triplet, whilst the lower-frequency outer sideband is enhanced and the higher-frequency one suppressed when $`\delta =\mathrm{\Omega }_R`$ and $`2\mathrm{\Omega }_R`$. These features are similar to those of a laser-dressed two-level atom coupled to such a cavity field . The enhancement and suppression of the sidebands is due to the cavity modification of the transition rates $`|a|c`$ and $`|c|a`$. The former decreases while the latter increases for $`\delta =\mathrm{\Omega }_R,\mathrm{\hspace{0.17em}2}\mathrm{\Omega }_R`$. Therefore, the enhanced spectral line of the outer sidebands is narrowed whilst the suppressed one is broadened. For large values of the detuning, shown in frame (d), the spectrum reverts to a symmetric form.
Next, we display the modified spectrum in the limit of $`\omega _{21}\mathrm{\Omega }`$ in Fig. 6 where the values $`\omega _{21}=200`$ and $`\mathrm{\Omega }=50`$ are taken. Contrary to the spectra of Fig. 5, the inner sidebands are most pronounced for the $`\delta =0`$ situation whereas the outer spectral lines are almost invisible. Tuning the cavity frequency may also change the spectral profile: the higher-frequency sideband is somewhat enhanced when $`\delta =\mathrm{\Omega }_R`$ and $`2\mathrm{\Omega }_R`$. As we discussed above, in the limit of $`\omega _{21}\mathrm{\Omega }`$ the dressed state populations $`\rho _{aa}`$ and $`\rho _{cc}`$ are close to zero while $`\rho _{bb}1`$, and all the populations are barely dependent on the cavity frequency. It is obvious from eqs. (72) and (73) that the $`\mathrm{\Lambda }_{1,2}`$ are mainly determined by the cavity-frequency-dependent decay rates, $`\mathrm{\Gamma }_{3a}`$ and $`\mathrm{\Gamma }_{3b}`$. For the parameters of Fig. 6, we have $`\mathrm{\Gamma }_{3a}(\delta =0)=\mathrm{\Gamma }_{3a}(\delta =0)2.51`$, $`\mathrm{\Gamma }_{3a}(\delta =\mathrm{\Omega }_R)1.33,\mathrm{\Gamma }_{3b}(\delta =\mathrm{\Omega }_R)4.40`$ and $`\mathrm{\Gamma }_{3a}(\delta =2\mathrm{\Omega }_R)=0.93,\mathrm{\Gamma }_{3b}(\delta =2\mathrm{\Omega }_R)2.50`$. Therefore, $`\mathrm{\Lambda }_1\left(\omega =\mathrm{\Omega }_R\right)<\mathrm{\Lambda }_2\left(\omega =\mathrm{\Omega }_R\right)`$, which implies that the lower-frequency peak is lower than the higher-frequency sideband.
In Fig. 7, we take the values $`\omega _{21}=200,\mathrm{\Omega }=100,`$ which corresponds to the populations shown in frame (d) of Fig. 4. For $`\delta =0,`$ the bulk of the population is concentrated in state $`|b,`$ and there is little population in states $`|a`$and $`|c.`$ Hence the outer sidebands are weak in amplitude, whilst the inner sidebands and the central peak are pronounced. Tuning the laser to $`\delta =\mathrm{\Omega }_R`$ or $`\delta =2\mathrm{\Omega }_R`$ increases the relative amplitude of the high frequency or low frequency inner sideband respectively. As $`\delta `$ is increased through $`\delta =\mathrm{\Omega }_R`$ to $`\delta =2\mathrm{\Omega }_R,`$ the population in state $`|a`$ increases, whilst that in state $`|c`$ decreases. The low frequency outer sideband consequently increases in amplitude at the expence of the amplitude of the high frequency outer sideband.
For the general case, by inspection of equations (68) for the cavity-induced decay rates $`\mathrm{\Gamma }_i`$, one finds in the limit $`\mathrm{\Omega }_R\kappa `$ that if the cavity is tuned to resonance with the driving laser $`(\delta =0)`$, the rates are the same as those in free space , except for $`\mathrm{\Gamma }_3`$ and $`\mathrm{\Gamma }_5`$ being replaced by $`\mathrm{\Gamma }_3+\gamma _c2\eta ^2`$ and $`\mathrm{\Gamma }_5+\gamma _c8\eta ^2`$ respectively. This reflects the fact that the cavity-induced spontaneous emission rates can be greatly suppressed by increasing the Rabi frequency . As shown in Fig. 8(a) for $`\omega _{21}=200`$ and $`\mathrm{\Omega }=200`$, the inner and outer sidebands are broadened, whilst the central peak is narrowed.
We illustrate the spectra for $`\delta =\mathrm{\Omega }_R`$ and $`2\mathrm{\Omega }_R`$ in frames 8(b) and 8(c), respectively, where asymmetric spectral features are exhibited. The previous explanations apply to this case as well, that is, the enhancement and suppression of the spectral lines stems from the cavity modification of the dressed state population distribution and decay rates. For example, for $`\delta =2\mathrm{\Omega }_R`$ the population in the dressed state $`|a`$ is much greater than those in the dressed states $`|b`$ and $`|c`$, as shown in Fig. 4(e). Accordingly, the lower-frequency peaks are higher than their counterparts in the high frequency side.
## V Absorption spectrum
It is also natural to study the absorption spectrum of a weak, tunable probe field transmitted through the system, as we would expect this to be largely determined by the dressed states populations. The frequency $`\omega `$ of the driving laser is kept fixed, and the absorption is measured as a function of the frequency $`\nu `$ of the probe laser, measured from the laser frequency, $`\omega .`$ The absorption spectrum is given by
$$๐(\nu )=Re\underset{0}{\overset{\mathrm{}}{}}\left\{[A_{20}(t+\tau ),A_{02}(t)]+[A_{10}(t+\tau ),A_{01}(t)]\right\}_t\mathrm{}e^{i\nu \tau }๐\tau .$$
(76)
We present only a brief discussion of this phenomenon here. As for the resonance fluorescence spectrum, we may use the quantum regression theorem to express the probe spectrum as
$$๐(\omega )=Re\left[\stackrel{~}{F}_{01}\left(z\right)+\stackrel{~}{G}_{02}\left(z\right)\stackrel{~}{F}_{10}^{}\left(z\right)\stackrel{~}{G}_{20}^{}\left(z\right)\right]_{z=i\nu }$$
(77)
where
$$F^{}\left(\tau \right)=U^{}\left(\tau \right)\left[\rho \left(\mathrm{}\right)|10|\right]U\left(\tau \right)\text{and }G^{}\left(\tau \right)=U^{}\left(\tau \right)\left[\rho \left(\mathrm{}\right)|20|\right]U\left(\tau \right).$$
(78)
It is possible to use the secular approximation to obtain expressions for the spectral components, as was done in the previous section. However, for brevity, we concentrate on a qualitative discussion here.
Figure 9, frames (a)โ(c), shows the spectra for $`\omega _{21}=10,\mathrm{\Omega }=100,(\mathrm{\Omega }_R=141.5)`$ when most of the population resides in the dressed states $`|a`$ and $`|c,`$ and the population in state $`|b`$ is practically zero, although the latter does show a very small maximum for zero detuning, $`\delta =0.`$ The case $`\delta =0`$ is shown in frame (a). Because there is no population in state $`|b,`$ the absorption spectrum is largely determined by the transitions between the outer two levels from $`|c`$ to $`|a,`$ at frequency $`\nu 2\mathrm{\Omega }_R,`$ and by the transitions between the inner two levels $`|a`$ to $`|c,`$ at frequency $`\nu 2\mathrm{\Omega }_R.`$ For example, from Figure 4, following the arguments of Cohen-Tannoudji and Reynaud , one can write down an approximate expression for the weight of the line at $`\nu 2\mathrm{\Omega }_R`$ as
$$w_a\left(2\mathrm{\Omega }_R\right)P_aR_{ac}P_cR_{ca},$$
(79)
a negative value for $`w`$ representing amplification, and a positive value absorption. It is easily seen that
$$w_a(+2\mathrm{\Omega }_R)P_cR_{ca}P_aR_{ac},=w_a(2\mathrm{\Omega }_R)$$
(80)
Since the populations in levels $`|a`$ and $`|c`$ are equal when $`\delta =0,`$ the difference between absorption and amplification is determined by the value of the rates.
A similar argument gives
$$w_a\left(\mathrm{\Omega }_R\right)P_cR_{cb}+P_bR_{ba}P_bR_{bc}P_aR_{ab}P_cR_{cb}P_aR_{ab}.$$
(81)
Using the expressions (44) for the rates, we find
$$\left|\frac{w_a\left(\mathrm{\Omega }_R\right)}{w_a\left(+2\mathrm{\Omega }_R\right)}\right|\frac{\omega _{21}^2}{4\mathrm{\Omega }_R^2}1.$$
(82)
These arguments explain the main features of the plot presented in frame (a). In frame (b), we tune the driving laser to $`\delta =\mathrm{\Omega }_R`$ which has the effect of significantly increasing the population in $`|c`$ and decreasing that in level $`|a.`$ The magnitudes of the absorption at $`\nu 2\mathrm{\Omega }_R,`$ and the emission at $`\nu 2\mathrm{\Omega }_R,`$ are thus greatly increased. The difference between the populations $`|a`$ and $`|c`$ reaches a maximum for $`\delta =\pm 2\mathrm{\Omega }_R`$ . This value of $`\delta `$ is assumed in frame (c), where the greatest absorption/emission occurs.
In frames (d)โ(f), we change the parameters to $`\omega _{21}=200,\mathrm{\Omega }=50\left(\mathrm{\Omega }_R=122.5\right).`$ The populations for this case, whose behaviour varies greatly from the previous case, are shown in frame (c) of Fig. 4. Now we have a large population in state $`|b,`$ and only small populations in states $`|a`$ and $`|c.`$ The absorption spectra are thus quite different: the strongest features occur at $`\nu =\pm \mathrm{\Omega }_R,`$ and have a strong dispersive element, whereas the features at $`\nu =\pm 2\mathrm{\Omega }_R`$ are relatively insignificant. Tuning the driving laser to $`\delta =\pm \mathrm{\Omega }_R`$ or $`\delta =\pm 2\mathrm{\Omega }_R`$ again has the effect of greatly increasing the strengths of the absorptions/emissions.
The first three frames of Fig. 10 show the absorption spectra for $`\omega _{21}=200,\mathrm{\Omega }=100\left(\mathrm{\Omega }_R=173.2\right)`$ and the second three frames the absorption spectra for $`\omega _{21}=200,\mathrm{\Omega }=200\left(\mathrm{\Omega }_R=300\right).`$ In these two cases, changing the detuning $`\delta `$ has a greater effect than in the cases of the previous figure. This is particularly true in the last three frames, where it can be seen that changing the detuning can cause a switch from strong absorption to strong emission, and vice-versa.
## VI Conclusions
We have investigated, in the bad cavity limit, the resonance fluorescence of a V-type three-level atom strongly driven by a laser field and weakly coupled to a cavity mode, and we have demonstrated how the fluorescence may be controlled and manipulated by varying the cavity and Rabi frequencies. A strong dependency of the bare and dressed state populations on the cavity resonant frequency is shown. Population inversion in both the bare and the dressed-state bases can be achieved for appropriate atom-cavity coupling constants, cavity resonant frequency and high driving intensities. These population inversions result from the enhanced atom-cavity interaction when the cavity is tuned to resonance with the atomic dressed-state transition. The resultant fluorescence spectrum is also strongly dependent on the cavity frequency. When the cavity is in resonant with the driving field, the spectrum is symmetric. Specifically, if the excited levels of the atom are degenerate (or near-degenerate), a Mollow-like triplet is exhibited with central line narrowing, but the spectrum has a two-peak structure if the level splitting is much greater than the laser intensity. Otherwise, the spectrum consists of five peaks. When the cavity is tuned to resonance with one of the spectral sidebands, some of the spectral lines may be enhanced and others suppressed, and so the spectrum is asymmetric. Dynamical line-narrowing and peak-suppression of certain spectral lines can be achieved by increasing the laser intensity. We have also demonstrated that a large degree of control and manipulation of the probe absorption spectrum can be achieved by tuning the cavity.
###### Acknowledgements.
JSP wishes to thank the Royal Society London for the financial support. This work is supported by the Natural Science Foundation of China, and the United Kingdom EPSRC. |
warning/0003/math0003107.html | ar5iv | text | # Variations on deformation quantization
## 1 Introduction
Quantization of a classical system is a way to pass from classical to quantum results.
Classical mechanics, in its Hamiltonian formulation on the motion space, has for framework a symplectic manifold (or more generally a Poisson manifold). Observables are families of smooth functions on that manifold $`M`$. The dynamics is defined in terms of a Hamiltonian $`HC^{\mathrm{}}(M)`$ and the time evolution of an observable $`f_tC^{\mathrm{}}(M\times )`$ is governed by the equation :
$$\frac{d}{dt}f_t=\{H,f_t\}.$$
Quantum mechanics, in its usual Heisenbergโs formulation, has for framework a Hilbert space (states are rays in that space). Observables are families of selfadjoint operators on the Hilbert space. The dynamics is defined in terms of a Hamiltonian $`H`$, which is a selfadjoint operator, and the time evolution of an observable $`A_t`$ is governed by the equation :
$$\frac{dA_t}{dt}=\frac{i}{\mathrm{}}[H,A_t].$$
A natural suggestion for quantization is a correspondence $`Q:fQ(f)`$ mapping a function $`f`$ to a self adjoint operator $`Q(f)`$ on a Hilbert space $``$ in such a way that $`Q(1)=Id`$ and
$$[Q(f),Q(g)]=i\mathrm{}Q(\{f,g\}).$$
There is no such correspondence defined on all smooth functions on $`M`$ when one puts an irreducibility requirement which is necessary not to violate Heisenbergโs principle.
Different mathematical treatments of quantization appeared to deal with this problem:
* Geometric Quantization of Kostant and Souriau. This proceeds in two steps; first prequantization of a symplectic manifold $`(M,\omega )`$ where one builds a Hilbert space and a correspondence $`Q`$ as above defined on all smooth functions on $`M`$ but with no irreducibility, then polarization to โcut down the number of variablesโ. One succeeds to quantize only a small class of functions.
* Berezinโs quantization where one builds on a particular class of Kรคhler manifolds a family of associative algebras using a symbolic calculus, i.e. a dequantization procedure.
* Deformation Quantization introduced by Flato, Lichnerowicz and Sternheimer in and developed in where they
โ suggest that quantization be understood as a deformation of the structure of the algebra of classical observables rather than a radical change in the nature of the observables.โ
This deformation approach to quantization is part of a general deformation approach to physics. This was one of the seminal ideas stressed by Moshe: one looks at some level of a theory in physics as a deformation of another level .
Deformation quantization is defined in terms of a star product which is a formal deformation of the algebraic structure of the space of smooth functions on a Poisson manifold. The associative structure given by the usual product of functions and the Lie structure given by the Poisson bracket are simultaneously deformed.
The plan of this presentation is the following :
* Examples and existence of star products. After some definitions, I recall the history of the proofs of existence and through the history of some constructions of star products, starting from the Moyal star product on a vector space endowed with a constant Poisson structure. I give the standard star product on the dual of a Lie algebra, and recall Kontsevich star product for any Poisson structure on a vector space. I prove by elementary methods that all universal star products on the dual of a Lie algebra are essentially equivalent.
* Equivalence of star products. I first wander through the history of the parametrization of equivalence classes of star products. I define Deligneโs cohomology classes associated to differential star products on symplectic manifolds and show by ฤech methods how this yields an intrinsic parametrization of the equivalence classes of differential star products and how this allows to study automorphisms of a star product (in the symplectic framework). Finally, I define a generalised moment map for star products.
* Convergence of star products. I study mostly the convergence of a Berezin type star product on Hermitian symmetric spaces. The construction of such a star product involves a correspondence between operators (on the Hilbert space given by geometric quantization) and functions (their Berezin symbols). A parameter is introduced in the construction (generalising the power of the line bundle in geometric quantization). Asymptotic expansion in this parameter on a large algebra of functions yields a deformed product. One proves associativity and convergence of this product.
I would like to thank my friends Georges Pinczon and Daniel Sternheimer who brought many improvements to this presentation.
## 2 Examples and existence of star products
###### Definition 1
A Poisson bracket defined on the space of smooth functions on a manifold $`M`$, is a $``$\- bilinear map on $`C^{\mathrm{}}(M)`$, $`(u,v)\{u,v\}`$ such that for any $`u,v,wC^{\mathrm{}}(M)`$:
\- $`\{u,v\}=\{v,u\}`$;
\- $`\{\{u,v\},w\}+\{\{v,w\},u\}+\{\{w,u\},v\}=0`$;
\- $`\{u,vw\}=\{u,v\}w+\{u,w\}v`$.
A Poisson bracket is given in terms of a contravariant skew symmetric 2-tensor $`P`$ on $`M`$, called the Poisson tensor, by
$$\{u,v\}=P(dudv).$$
The Jacobi identity for the Poisson bracket Lie algebra is equivalent to the vanishing of the Schouten bracket :
$$[P,P]=0.$$
(The Schouten bracket is the extension -as a graded derivation for the exterior product- of the bracket of vector fields to skewsymmetric contravariant tensor fields.)
A Poisson manifold, denoted $`(M,P)`$, is a manifold $`M`$ with a Poisson bracket defined by the Poisson tensor $`P`$.
A particular class of Poisson manifolds, essential in classical mechanics, is the class of symplectic manifolds. If $`(M,\omega )`$ is a symplectic manifold (i.e. $`\omega `$ is a closed nondegenerate $`2`$-form on $`M`$) and if $`u,vC^{\mathrm{}}(M)`$, the Poisson bracket of $`u`$ and $`v`$ is defined by
$$\{u,v\}:=X_u(v)=\omega (X_v,X_u),$$
where $`X_u`$ denotes the Hamiltonian vector field corresponding to the function $`u`$, i.e. such that $`i(X_u)\omega =du.`$ In coordinates the components of the corresponding Poisson tensor $`P^{ij}`$ form the inverse matrix of the components $`\omega _{ij}`$ of $`\omega `$.
Duals of Lie algebras form the class of linear Poisson manifolds. If $`๐ค`$ is a Lie algebra then its dual $`๐ค^{}`$ is endowed with the Poisson tensor $`P`$ defined by
$$P_\xi (X,Y):=\xi ([X,Y])$$
for $`X,Y๐ค=(๐ค^{})^{}(T_\xi ๐ค^{})^{}`$.
###### Definition 2
(Bayen et al. ) A star product on $`(M,P)`$ is a bilinear map
$$N\times NN[[\nu ]],(u,v)uv=u_\nu v:=\underset{r0}{}\nu ^rC_r(u,v)$$
where $`N=C^{\mathrm{}}(M)`$, such that
1. when the map is extended $`\nu `$-linearly (and continuously in the $`\nu `$-adic topology) to $`N[[\nu ]]\times N[[\nu ]]`$ it is formally associative:
$$(uv)w=u(vw);$$
2. (a) $`C_0(u,v)=uv`$, (b) $`C_1(u,v)C_1(v,u)=\{u,v\}`$;
3. $`1u=u1=u`$.
When the $`C_r`$โs are bidifferential operators on $`M`$, one speaks of a differential star product.
###### Remark 3
A star product can also be defined not on the whole of $`C^{\mathrm{}}(M)`$ but on any subspace $`N`$ of it which is stable under pointwize multiplication and Poisson bracket.
Requiring differentiability of the cochains is essentially the same as requiring them to be local .
In (b) we follow Deligneโs normalisation for $`C_1`$: its skew symmetric part is $`\frac{1}{2}\{,\}`$. In the original definition it was equal to the Poisson bracket. One finds in the literature other normalisations such as $`\frac{i}{2}\{,\}`$. All these amount to a rescaling of the parameter $`\nu `$.
One assumed also the parity condition $`C_r(u,v)=(1)^rC_r(v,u)`$ in the earliest definition.
Property (b) above implies that the centre of $`C^{\mathrm{}}(M)[[\nu ]]`$, when the latter is viewed as an algebra with multiplication $``$, is a series whose terms Poisson commute with all functions, so is an element of $`[[\nu ]]`$ when $`M`$ is symplectic and connected.
Properties (a) and (b) of Definition 2 imply that the star commutator defined by $`[u,v]_{}=uvvu`$, which obviously makes $`C^{\mathrm{}}(M)[[\nu ]]`$ into a Lie algebra, has the form $`[u,v]_{}=\nu \{u,v\}+\mathrm{}`$ so that repeated bracketing leads to higher and higher order terms. This makes $`C^{\mathrm{}}(M)[[\nu ]]`$ an example of a pronilpotent Lie algebra. We denote the star adjoint representation $`ad_{}u(v)=[u,v]_{}.`$
### 2.1 The Moyal star product on $`^n`$
The simplest example of a deformation quantization is the Moyal product for the Poisson structure $`P`$ on a vector space $`V=^m`$ with constant coefficients:
$$P=\underset{i,j}{}P^{ij}_i_j,P^{ij}=P^{ji}$$
where $`_i=/x^i`$ is the partial derivative in the direction of the coordinate $`x^i,i=1,\mathrm{},n`$. The formula for the Moyal product is
$$(u_Mv)(z)=\mathrm{exp}\left(\frac{\nu }{2}P^{rs}_{x^r}_{y^s}\right)(u(x)v(y))|_{x=y=z}.$$
(1)
When $`P`$ is non degenerate (so $`V=^{2n}`$), the space of formal power series of polynomials on $`V`$ with Moyal product is called the Weyl algebra $`W=(S(V)[[\nu ]],_M)`$.
This example comes from the composition of operators via Weylโs quantization. Weylโs correspondence associates to certain functions $`f`$ on $`^{2n}`$ an operator $`W(f)`$ on $`L^2(^n)`$
$$W(f)=\stackrel{~}{f}(\xi ,\eta )\mathrm{exp}\left(\frac{i(\xi P+\eta Q)}{\mathrm{}}\right)๐\xi ๐\eta $$
where $`\stackrel{~}{f}`$ is the inverse Fourier transform of $`f`$. Then
$$W(f)W(g)=W(f_Mg)(\nu =i\mathrm{}).$$
In fact, Moyal had used in 1949 the deformed bracket which corresponds to the commutator of operators to study quantum statistical mechanics. The Moyal product first appeared in Groenewold.
$``$ In 1974, Flato, Lichnerowicz and Sternheimer studied deformations of the Lie algebra structure defined by the Poisson bracket on the algebra $`N`$ of smooth functions on a symplectic manifold; they studied 1-differential deformations because the relevant cohomology, i.e. the 1-differential Chevalley cohomology of the Lie algebra $`N`$ with values in $`N`$ for the adjoint representation, was known .
In 1975, Vey pursued their work in the differential context: he constructed a differential deformation on $`M=^{2n}`$ which turns out to be the Moyal bracket $`\{u,v\}_M:=\frac{1}{\nu }(u_Mvv_Mu)`$ and he proved that there exists a differential deformation on a symplectic manifold when its third de Rham cohomology space is trivial ($`b_3(M):=dimH^3(M;)=0`$). He also reconstructed the Moyal product on $`^{2n}`$.
This opened the path to deformation quantization presented in 1976 by Flato, Lichnerowicz and Sternheimer .
$``$ In 1978, in their seminal paper about deformation quantization , Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer proved that Moyal star product can be defined on any symplectic manifold $`(M,\omega )`$ which admits a symplectic connection $``$ (i.e. a linear connection such that $`\omega =0`$ and the torsion of $``$ vanishes) with no curvature. They also built star products on some quotient of $`^{2n}`$.
$``$ In 1979, Neroslavsky and Vlassov proved with Lichnerowicz that on any symplectic manifold with $`b_3(M)=0`$, there exists a differential star product.
### 2.2 Hochschild cohomology
The study of star products on a manifold $`M`$ used Gerstenhaber theory of deformations of associative algebras. This uses the Hochschild cohomology of the algebra, here $`C^{\mathrm{}}(M)`$ with values in $`C^{\mathrm{}}(M)`$, where $`p`$-cochains are $`p`$-linear maps from $`(C^{\mathrm{}}(M))^p`$ to $`C^{\mathrm{}}(M)`$ and where the Hochschild coboundary operator maps the $`p`$-cochain $`C`$ to the $`p+1`$-cochain
$`(C)(u_0,\mathrm{},u_p)=u_0C(u_1,\mathrm{},u_p)`$
$`+{\displaystyle \underset{r=1}{\overset{p}{}}}(1)^rC(u_0,\mathrm{},u_{r1}u_r,\mathrm{},u_p)+(1)^{p+1}C(u_0,\mathrm{},u_{p1})u_p.`$
For differential star products, we consider differential cochains, i.e. given by differential operators on each argument. The associativity condition for a star product at order $`k`$ in the parameter $`\nu `$ reads
$$(C_k)(u,v,w)=\underset{r+s=k,r,s>0}{}\left(C_r(C_s(u,v),w)C_r(u,C_s(v,w))\right).$$
If one has cochains $`C_j,j<k`$ such that the star product they define is associative to order $`k1`$, then the right hand side above is a cocycle ($``$(RHS)$`=0`$) and one can extend the star product to order $`k`$ if it is a coboundary (RHS$`=(C_k))`$.
Denoting by $`m`$ the usual multiplication of functions, and writing $`=m+C`$ where $`C`$ is a formal series of multidifferential operators ($`CD_{poly}(M)[[\nu ]]`$) the associativity also reads $`C=[C,C]`$ where the bracket on the right hand side is the graded Lie algebra bracket on $`D_{poly}(M)[[\nu ]]`$.
###### Theorem 4
(Vey ) Every differential $`p`$-cocycle $`C`$ on a manifold $`M`$ is the sum of the coboundary of a differential (p-1)-cochain and a $`1`$-differential skewsymmetric $`p`$-cocycle $`A`$:
$$C=B+A.$$
In particular, a cocycle is a coboundary if and only if its total skewsymmetrization, which is automatically $`1`$-differential in each argument, vanishes. Also
$$H_{\mathrm{diff}}^p(C^{\mathrm{}}(M),C^{\mathrm{}}(M))=\mathrm{\Gamma }(\mathrm{\Lambda }^pTM).$$
Furthermore (),given a connection $``$ on $`M`$, $`B`$ can be defined from $`C`$ by universal formulas.
By universal, we mean the following: any $`p`$-differential operator $`D`$ of order maximum $`k`$ in each argument can be written
$$D(u_1,\mathrm{},u_p)=\underset{|\alpha _1|<k\mathrm{}|\alpha _p|<k}{}D_{|\alpha _1|,\mathrm{},|\alpha _p|}^{\alpha _1\mathrm{}\alpha _p}_{\alpha _1}u_1\mathrm{}_{\alpha _p}u_p$$
where $`\alpha `$โs are multiindices, $`D_{|\alpha _1|,\mathrm{},|\alpha _p|}`$ are tensors (symmetric in each of the $`p`$ groups of indices) and $`_\alpha u=(\mathrm{}(u))(\frac{}{x^{i_1}},\mathrm{},\frac{}{x^{i_q}})`$ when $`\alpha =(i_1,\mathrm{},i_q)`$. We claim that there is a $`B`$ such that the tensors defining $`B`$ are universally defined as linear combinations of the tensors defining $`C`$, universally meaning in a way which is independent of the form of $`C`$. Note that requiring differentiability of the cochains is essentially the same as requiring them to be local .
(An elementary proof of the above theorem can be found in .)
###### Remark 5
Behind theorem 4 above, are the following stronger results about Hochschild cohomology:
###### Theorem 6
Let $`๐=C^{\mathrm{}}(M)`$, let $`๐(๐)`$ be the space of continuous cochains and $`๐_{diff}(๐)`$ be the space of differential cochains. Then
* $`\mathrm{\Gamma }(\mathrm{\Lambda }^pTM)H^p(C^{\mathrm{}}(M),C^{\mathrm{}}(M))`$;
* the inclusions $`\mathrm{\Gamma }(\mathrm{\Lambda }^pTM)๐_{diff}(๐)๐(๐)`$ induce isomorphisms in cohomology.
Point 1 follows from the fact that any cochain which is 1-differential in each argument is a cocycle and that the skewsymmetric part of a coboundary always vanishes. The fact that the inclusion $`\mathrm{\Gamma }(\mathrm{\Lambda }TM)๐_{diff}(๐)`$ induces an isomorphism in cohomology is proven by Vey ; it gives theorem 4. The general result about continuous cochains is due to Connes . Another proof of Connes result was given by Nadaud in . In the somewhat pathological case of completely general cochains the full cohomology does not seem to be known.
$``$ Some examples of star products were built using inductively the explicit formulas for coboundaries on locally symmetric symplectic manifolds in ; the cochains are given by universal expressions in the symplectic $`2`$-form, its inverse (i.e. the Poisson tensor) and the curvature tensor of the symmetric symplectic connexion.
$``$ We also showed that assuming a homogeneity condition on the cochains proves the existence of a star product on the cotangent bundle to any parallelisable manifold and gave explicit formulas for the cotangent bundle to a Lie group. The vertical part of this gives a deformation quantization of any linear Poisson manifold, i.e. any dual of a Lie algebra .
### 2.3 The standard \*-product on $`๐ค^{}`$
Let $`๐ค^{}`$ be the dual of a Lie algebra $`๐ค`$. The algebra of polynomials on $`๐ค^{}`$ is identified with the symmetric algebra $`S(๐ค)`$. One defines a new associative law on this algebra by a transfer of the product $``$ in the universal enveloping algebra $`U(๐ค)`$, via the bijection between $`S(๐ค)`$ and $`U(๐ค)`$ given by the total symmetrization $`\sigma `$ :
$$\sigma :S(๐ค)U(๐ค)X_1\mathrm{}X_k\frac{1}{k!}\underset{\rho S_k}{}X_{\rho (1)}\mathrm{}X_{\rho (k)}.$$
Then $`U(๐ค)=_{n0}U_n`$ where $`U_n:=\sigma (S^n(๐ค))`$ and we decompose an element $`uU(๐ค)`$ accordingly $`u=u_n`$. We define for $`PS^p(๐ค)`$ and $`QS^q(๐ค)`$
$$PQ=\underset{n0}{}(\nu )^n\sigma ^1((\sigma (P)\sigma (Q))_{p+qn}).$$
(2)
This yields a differential star product on $`๐ค^{}`$ . Using Vergneโs result on the multiplication in $`U(๐ค)`$, this star product is characterised by
$`XX_1\mathrm{}X_k=XX_1\mathrm{}X_k`$
$`+{\displaystyle \underset{j=1}{\overset{k}{}}}{\displaystyle \frac{(1)^j}{j!}}\nu ^jBj[[[X,X_{r_1}],\mathrm{}],X_{r_j}]X_1\mathrm{}\widehat{X_{r_1}}\mathrm{}\widehat{X_{r_j}}\mathrm{}X_k`$
where $`B_j`$ are the Bernouilli numbers. This star product can be written with an integral formula (for $`\nu =2\pi i`$) :
$$uv(\xi )=_{๐ค\times ๐ค}\widehat{u}(X)\widehat{v}(Y)e^{2i\pi \xi ,CBH(X,Y)}๐X๐Y$$
where $`\widehat{u}(X)=_๐ค^{}u(\eta )e^{2i\pi \eta ,X}`$ and where $`CBH`$ denotes Campbell-Baker-Hausdorff formula for the product of elements in the group in a logarithmic chart ($`\mathrm{exp}X\mathrm{exp}Y=\mathrm{exp}CBH(X,Y)X,Y๐ค`$).
We call this the standard (or CBH) star product on the dual of a Lie algebra.
$``$ An important property of this star product is its covariance, i.e. the fact that
$$XYYX=\nu [X,Y]X,Y๐ค.$$
In general, when there is a Hamiltonian action of a connected Lie group $`G`$ on a manifold $`M`$, denoting by $`\lambda _X`$ the function on $`M`$ corresponding to $`X๐ค`$, a star product on $`M`$ is covariant under $`G`$ if $`\lambda _X\lambda _Y\lambda _Y\lambda _X=\nu \lambda _{\{X,Y\}}X,Y๐ค`$. Arnal, Cortet, Molin and Pinczon have proven in that for any covariant star product there is a representation of $`G`$ into the automorphisms of the star product. We come back to this notion of covariance in ยง3.2.
A program to study representations through deformation quantization methods was introduced in 1978 by Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer in and in Fronsdal . The notion of covariant star products is essential for this. The program involves a beautiful notion of adapted Fourier transform (or $``$ exponential). I shall not speak about this aspect of deformation quantization; let me just mention that the program to study representations through deformations was carried first in the case of nilpotent Lie groups in the mid eighties by Arnal and Cortet and that the adapted Fourier transform they found gives a Kirillov type formula for the character of the representation and a localisation of the (packets of) coefficients of representations on the corresponding orbits in the dual of the Lie algebra. This theory of star representations was developed for all representations of compact groups and exponential groups, and for some series of representations of semi-simple groups.
$``$ The standard star product on $`๐ค^{}`$ does not restrict to orbits (except for the Heisenberg group) so other algebraic constructions of star products on $`S(๐ค)`$ were considered (with Michel Cahen in , with Cahen and Arnal in , by Arnal, Ludwig and Masmoudi in and more recently by Fioresi and Lledo in ). For instance, when $`๐ค`$ is semisimple, if $``$ is the space of harmonic polynomials and if $`I_1,\mathrm{}I_r`$ are generators of the space of invariant polynomials, then any polynomial $`PS(๐ค)`$ writes uniquely as a sum $`P=_{a_1\mathrm{}a_r}I_1^{a_1}\mathrm{}I_r^{a_r}h_{a_1\mathrm{}a_r}`$ where $`h_{a_1\mathrm{}a_r}`$. One considers the isomorphism $`\sigma ^{}`$ between $`S(๐ค)`$ and $`U(๐ค)`$ induced by this decomposition
$$\sigma ^{}(P)=\underset{a_1\mathrm{}a_r}{}(\sigma (I_1))^{a_1}\mathrm{}(\sigma (I_r))^{a_r}\sigma (h_{a_1\mathrm{}a_r}).$$
This gives a star product on $`S(๐ค)`$ which is not defined by differential operators. In fact, with Cahen and Rawnsley, we proved that if $`๐ค`$ is semisimple, there is no differential star product on any neighbourhood of $`0`$ in $`๐ค^{}`$ such that $`Cu=Cu`$ for the quadratic invariant polynomial $`CS(๐ค)`$ and $`uS(๐ค)`$ (thus no differential star product which is tangential to the orbits).
$``$ In 1983, De Wilde and Lecomte proved that on any symplectic manifold there exists a differential star product. This was obtained by imagining a very clever generalisation of a homogeneity condition in the form of building at the same time the star product and a special derivation of it. A very nice presentation of this proof appears in . Their technique works to prove the existence of a differential star product on a regular Poisson manifold .
$``$ In 1985, but appearing only in the West in the nineties , Fedosov gave a recursive construction of a star product on a symplectic manifold $`(M,\omega )`$ by
-considering the Weyl bundle $`๐ฒ`$ which is the bundle on $`M`$ in associative Weyl algebras $`W`$ associated to the principal bundle of symplectic frames;
-building, from a symplectic connection $``$ on $`M`$, a covariant derivative on $`\mathrm{\Gamma }(๐ฒ)`$, $`=^{}+[r,.](r\mathrm{\Gamma }(T^{}M๐ฒ))`$, such that $`=0`$;
-identifying $`C^{\mathrm{}}(M)[[\nu ]]`$ with $`\{s\mathrm{\Gamma }(๐ฒ)s=0\}`$ and transferring the associative pointwize product of sections to an associative product on $`C^{\mathrm{}}(M)[[\nu ]]`$.
In 1994, he extended this result to give a recursive construction in the context of regular Poisson manifold .
$``$ Independently, also using the framework of Weyl bundles, Omori, Maeda and Yoshioka gave an alternative proof of existence of a differential star product on a symplectic manifold, gluing local Moyal star products.
$``$ In 1986, Drinfeld proposed a program of quantization of Poisson-Lie groups.
One way to consider the problem is to study deformations of the Hopf algebra $`U(๐ค)`$ of a finite dimensional Lie bialgebra $`(๐ค,p)`$, where one deforms the product and the coproduct, the deformation of the coproduct being driven by the cocycle $`p`$. By duality this is the quantization of formal Poisson-Lie groups. For the standard structures on semisimple Lie algebras, this was solved by Drinfeld; it is the construction of classical quantum groups. Drinfeld showed that the quantization is preferred, i.e. the product in $`U(๐ค)`$ is unchanged (or, by duality, the coproduct on formal series โ which are the functions on the formal Poisson-Lie groupโ is unchanged). Etingof and Kazhdan proved that one can deform $`U(๐ค)`$ for any Lie bialgebra $`(๐ค,p)`$, so one can always quantize formal Poisson-Lie groups. They proved that the deformation is differential. Pinczon proved that any preferred quantization of a formal Poisson-Lie group is differential.
Quantum groups became very popular and had fundamental applications. It is not my goal to describe those developments but they brought new interest in the question of deformation quantization of general Poisson manifolds.
Indeed, another way to consider the problem of quantization of a Poisson-Lie group $`(G,P)`$ is to study deformations of the Hopf algebra $`C^{\mathrm{}}(G)`$, where the deformation of the product of functions is driven by $`P`$.
In the papers by Bonneau, Flato, Gerstenhaber and Pinczon and by Bidegain and Pinczon, the duality between those two approaches is proven in the framework of topological deformations. In particular, any classical quantum group gives, by topological duality, a differential deformation of $`C^{\mathrm{}}(G)`$. In , Bidegain and Pinczon have built a differential preferred deformation of $`C^{\mathrm{}}(G)`$ for any Poisson Lie group $`(G,P)`$ with $`G`$ semisimple. They also show that any preferred deformation of $`C^{\mathrm{}}(G)`$ is automatically differential. Then Etingof and Kazhdan proved that there exists a differential deformation of $`C^{\mathrm{}}(G)`$ for any Poisson Lie group $`(G,P)`$, deformation which is preferred in the quasi triangular case.
$``$ The existence of a star product on other classes of Poisson manifolds was studied by various authors (OmoriโMaedaโYoshioka , Tamarkin , Asin ,โฆ).
$``$ In 1997, Kontsevich gave a proof of the existence of a star product on any Poisson manifold and gave an explicit formula for a star product for any Poisson structure on $`V=^m`$. This appeared as a consequence of the proof of his formality theorem. Tamarkin gave a version of the proof in the framework of the theory of operads.
### 2.4 Kontsevich star product on $`V`$
Let $`M`$ be a domain in $`V=^m`$ and $`P`$ be any Poisson structure on $`M`$. Kontsevich builds a star product on $`(M,P)`$, where the star product of two functions $`u`$ and $`v`$ is given in terms of some universal polydifferential operators applied to the coefficients of the bi-vector field $`P`$ and to the functions $`u,v`$; the formula is invariant under affine transformations of $`^m`$ and the description of the $`k^{\mathrm{th}}`$ cochain uses a special class $`G_k`$ of oriented labelled graphs.
An (oriented) graph $`\mathrm{\Gamma }`$ is a pair $`(V_\mathrm{\Gamma },E_\mathrm{\Gamma })`$ of two finite sets such that $`E_\mathrm{\Gamma }`$ is a subset of $`V_\mathrm{\Gamma }\times V_\mathrm{\Gamma }`$; elements of $`V_\mathrm{\Gamma }`$ are vertices of $`\mathrm{\Gamma }`$, elements of $`E_\mathrm{\Gamma }`$ are edges of $`\mathrm{\Gamma }`$. If $`e=(v_1,v_2)E_\mathrm{\Gamma }V_\mathrm{\Gamma }\times V_\mathrm{\Gamma }`$ is an edge it is said that $`e`$ starts at $`v_1`$ and ends at $`v_2`$.
A graph $`\mathrm{\Gamma }`$ belongs to $`G_k`$ if $`\mathrm{\Gamma }`$ has $`k+2`$ vertices $`V_\mathrm{\Gamma }=\{1,\mathrm{},k\}\{L,R\}`$ and $`2k`$ labelled edges (with no multiple edges and no edge of the form $`(v,v)`$ for $`vV_\mathrm{\Gamma }`$), $`E_\mathrm{\Gamma }=\{e_1^1,e_1^2,e_2^1,e_2^2,\mathrm{},e_k^1,e_k^2\}`$ where $`e_j^1`$ and $`e_j^2`$ start at $`j`$.
To each labelled graph $`\mathrm{\Gamma }G_k`$ and each skew symmetric $`2`$-tensor $`P`$ is associated a bidifferential operator $`C_\mathrm{\Gamma }(P)`$ on $`M^m`$ :
$`(C_\mathrm{\Gamma }(P))(u,v):`$ $`=`$ $`{\displaystyle \underset{I:E_\mathrm{\Gamma }\{1,\mathrm{},m\}}{}}[{\displaystyle \underset{j=1}{\overset{k}{}}}({\displaystyle \underset{eE_\mathrm{\Gamma }e=(,j)}{}}_{I(e)})P^{I(e_j^1)I(e_j^2)}]`$
$`\times `$ $`({\displaystyle \underset{eE_\mathrm{\Gamma }e=(,L)}{}}_{I(e)})u\times ({\displaystyle \underset{eE_\mathrm{\Gamma }e=(,R)}{}}_{I(e)})v.`$
A weight $`w_\mathrm{\Gamma }`$ is associated with each graph $`\mathrm{\Gamma }G_k`$. Denote by $`_k`$ the space of configurations of $`k`$ numbered distinct points in the upper half plane $`=\{z|Im(z)>0\}`$:
$$_k=\{(p_1,\mathrm{},p_k)p_jp_ip_j\mathrm{for}ij\}.$$
For $`p,q`$ define
$$\varphi ^h(p,q)=\frac{1}{2i}Log\left(\frac{(qp)(\overline{q}p)}{(q\overline{p})(\overline{q}\overline{p})}\right).$$
If $`\mathrm{\Gamma }G_k`$ is a graph as above, assign a point $`p_j`$ to the vertex $`j`$ for $`1jk`$, point $`0`$ to the vertex $`L`$, and point $`1`$ to the vertex $`R`$. Every edge $`eE_\mathrm{\Gamma }`$ defines an ordered pair $`(p,q)`$ of points on $``$, thus a function $`\varphi _e^h=\varphi ^h(p,q)`$ on $`_k`$ with values in $`/2\pi `$.
The weight of $`\mathrm{\Gamma }`$ is defined by
$$w_\mathrm{\Gamma }=\frac{1}{k!(2\pi )^{2k}}__n_{i=1}^k(d\varphi _{e_i^1}^hd\varphi _{e_i^2}^h).$$
###### Theorem 7
(Kontsevich ) Let $`P`$ be a Poisson tensor on a domain of $`^m`$. Then
$$u_\nu v:=\underset{k=0}{\overset{\mathrm{}}{}}\nu ^k\underset{\mathrm{\Gamma }G_k}{}w_\mathrm{\Gamma }(C_\mathrm{\Gamma }(P))(u,v)$$
defines a differential star product.
###### Definition 8
A star product on $`^m`$ is universal if it is given, for any Poisson tensor $`P`$ by
$$u_\nu v=uv+\underset{n>0}{}\nu ^n(C_n(P))(u,v)\mathrm{with}\mathrm{C}_\mathrm{n}(\mathrm{P})=\underset{\mathrm{\Gamma }\mathrm{G}}{}\mathrm{w}^{}(\mathrm{\Gamma })\mathrm{C}_\mathrm{\Gamma }(\mathrm{P})$$
where the $`w^{}(\mathrm{\Gamma })`$ are scalars which are independent of $`P`$.
###### Proposition 9
Two universal star products $`_\nu `$ and $`_\nu ^{}`$ on $`๐ค^{}`$ are always equivalent modulo a change of parameter. This means that there exists a series $`T(\nu )=Id+_{r=1}^{\mathrm{}}\nu ^rT_r`$ of linear universal differential operators and a series $`\nu ^{}=\nu +_{n>1}f_n\nu ^n`$ such that
$$u_\nu ^{}^{}v=T^1(\nu )(T(\nu )u_\nu T(\nu )v).$$
This was first proven by Arnal ; we give here an elementary proof of this fact.
###### Proof.
Assume by induction that $``$ and $`^{}`$ coincide at order $`n1`$. Associativity relation at order $`n`$ implies $`C_n(P)=C_n^{}(P)`$; thus $`C_n(P)=C_n^{}(P)+E_n(P)+A_n(P)`$ where $`E_n`$ is a universal differential operator and $`A_n(P)`$ is universal, skewsymmetric and of order $`(1,1)`$. Clearly, if $`A_n`$ corresponds to a graph in $`G_k`$, it has $`2k`$ arrows and $`k+2`$ vertices, but since $`P`$ is linear on $`๐ค^{}`$, at most one arrow can end at any of the first $`k`$ vertices and since $`A_n`$ is $`1`$-differential in each argument, exactly one arrow ends at each of the last $`2`$ vertices. Hence $`2kk+2`$ so $`k2`$; but if $`k=1`$ the only graph yields the Poisson bracket of functions so this can be cancelled by a change of parameter and if $`k=2`$ the graph corresponds to a symmetric bidifferential operator, hence does not yield a $`A_n(P)`$. So we can assume that $`A_n`$ vanishes but then the equivalence through $`T(\nu )=Id+\nu ^nE_n(P)`$ builds a star product which is universal and coincide with $``$ at order $`n`$. โ
###### Remark 10
When a covariance condition or a homogeneity condition ($`C_n(sP)=s^nC_n(P)`$) is added, clearly there is no need for a change of parameter since the graph in $`G_1`$ can only arise in $`C_1`$ so two such universal star products are always equivalent.
In particular, the star product of Kontsevich and the standard (CBH) star product on $`๐ค^{}`$, which in general are not the same, are equivalent. The equivalence is given by universal differential operators, hence combinations of wheels which are graphs consisting of $`k+1`$ points $`1,\mathrm{},k`$ and $`L`$ and $`2k`$ arrows $`(1,2),(2,3)\mathrm{},(k1,k),(k,1)`$ and $`(j,L)\mathrm{\hspace{0.17em}1}jk`$. Such a wheel clearly vanishes when the Poisson structure corresponds to a nilpotent Lie algebra so there is only one covariant universal star product on the dual of a nilpotent Lie algebra (this appears in Arnal and Kathotia ).
The equivalence between Kontsevich and CBH star product has been explicitly constructed (see Arnal and Dito ) and gives an integral formula for Kontsevich star product:
$$uv(\xi )=\stackrel{~}{u}(X)\stackrel{~}{v}(Y)\frac{F(X)F(Y)}{F(CBH(X,Y))}\mathrm{exp}(2\pi i\xi ,CBH(X,Y))๐X๐Y$$
where $`F`$ is some formal function on $`๐ค`$ written as a sum of products of traces of powers of $`ad(X)`$.
This star product has been used recently by Andler, Dvorsky and Sahi to establish a conjecture of Kashiwara and Vergne which, in turn, gives a new proof of Dufloโs result on the local solvability of bi-invariant differential operators on a Lie group.
## 3 Equivalence of star products
###### Definition 11
Two star products $``$ and $`^{}`$ on $`(M,P)`$ are said to be equivalent if there is a series
$$T=Id+\underset{r=1}{\overset{\mathrm{}}{}}\nu ^rT_r$$
where the $`T_r`$ are linear operators on $`C^{\mathrm{}}(M)`$, such that
$$T(fg)=Tf^{}Tg.$$
(3)
Remark that the $`T_r`$ automatically vanish on constants since $`1`$ is a unit for $``$ and for $`^{}`$. Using in a similar way linear operators which do not necessarily vanish on constants, one can pass from any associative deformation of the product of functions on a Poisson manifold $`(M,P)`$ to another such deformation with $`1`$ being a unit.
###### Definition 12
A Poisson deformation of the Poisson bracket on a Poisson manifold $`(M,P)`$ is a Lie algebra deformation of $`(C^{\mathrm{}}(M),\{,\})`$ which is a derivation in each argument, i.e. of the form $`\{u,v\}_\nu =P_\nu (du,dv)`$ where $`P_\nu =P+\nu ^kP_k`$ is a series of skewsymmetric contravariant $`2`$-tensors on $`M`$ (such that $`[P_\nu ,P_\nu ]=0`$).
Two Poisson deformations $`P_\nu `$ and $`P_\nu ^{}`$ of the Poisson bracket $`P`$ on a Poisson manifold $`(M,P)`$ are equivalent if there exists a formal path in the diffeomorphism group of $`M`$, starting at the identity, i. e. a series $`T=\mathrm{exp}D=Id+_j\frac{1}{j!}D^j`$ for $`D=_{r1}\nu ^rD_r`$ where the $`D_r`$ are vector fields on $`M`$, such that
$$T\{u,v\}_\nu =\{Tu,Tv\}_\nu ^{}$$
where $`\{u,v\}_\nu =P_\nu (du,dv)`$ and $`\{u,v\}_\nu ^{}=P_\nu ^{}(du,dv)`$.
$``$ In the general theory of deformations, Gerstenhaber showed how equivalence is linked to some second cohomology space.
$``$ For symplectic manifolds, Flato, Lichnerowicz and Sternheimer in 1974 studied $`1`$-differential deformations of the Poisson bracket ; it follows from their work, and appears in Lecomte , that:
###### Proposition 13
The equivalence classes of Poisson deformations of the Poisson bracket $`P`$ on a symplectic manifold $`(M,\omega )`$ are parametrised by $`H^2(M;)[[\nu ]]`$.
Indeed, one first show that any Poisson deformation $`P_\nu `$ of the Poisson bracket $`P`$ on a symplectic manifold $`(M,\omega )`$ is of the form $`P^\mathrm{\Omega }`$ for a series $`\mathrm{\Omega }=\omega +_{k1}\nu ^k\omega _k`$ where the $`\omega _k`$ are closed $`2`$-forms, and $`P^\mathrm{\Omega }(du,dv)=\mathrm{\Omega }(X_u^\mathrm{\Omega },X_v^\mathrm{\Omega })`$ where $`X_u^\mathrm{\Omega }=X_u+\nu (\mathrm{})\mathrm{\Gamma }(TM)[[\nu ]]`$ is the element defined by $`i(X_u^\mathrm{\Omega })\mathrm{\Omega }=du`$.
* Observe that for any series $`\mathrm{\Omega }`$ as above the series of $`1`$-differential $`2`$-cochains $`P^\mathrm{\Omega }`$ satisfies $`[P^\mathrm{\Omega },P^\mathrm{\Omega }]=0`$ because $`\mathrm{\Omega }`$ is a closed $`2`$-form, so defines indeed a Poisson deformation of $`P`$.
Reciprocally, given a $`1`$-differential deformation $`P_\nu =P+_j\nu ^jP_j`$, assume it coincides up to order $`k`$ ($`k0`$) with $`P^{\mathrm{\Omega }_k}`$ for some $`\mathrm{\Omega }_k=\omega +_{j1}^k\nu ^j\omega _j`$ then $`[P_\nu ,P_\nu ][P^{\mathrm{\Omega }_k},P^{\mathrm{\Omega }_k}]=0`$ imply at order $`k+1`$ that $`[P,P_{k+1}(P^{\mathrm{\Omega }_k})_{k+1}=0]`$ so that there exists a closed $`2`$-form $`\omega _{k+1}`$ such that $`(P_{k+1}(P^{\mathrm{\Omega }_k})_{k+1})(du,dv)=\omega _{k+1}(X_u,X_v)`$. Hence $`P_\nu `$ coincides up to order $`k+1`$ with $`P^{\mathrm{\Omega }_{k+1}}`$ where $`\mathrm{\Omega }_{k+1}=\omega +_{j1}^{k+1}\nu ^j\omega _j`$. By induction, any Poisson deformation of $`P`$ is of the form $`P^\mathrm{\Omega }`$ for a series $`\mathrm{\Omega }=\omega +_{k1}\nu ^k\omega _k`$ where the $`\omega _k`$ are closed $`2`$-forms.
One then shows that two Poisson deformations $`P^\mathrm{\Omega }`$ and $`P^\mathrm{\Omega }^{}`$ are equivalent if and only if $`\omega _k`$ and $`\omega _k^{}`$ are cohomologous for all $`k1`$.
* If $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ coincide to order $`k1`$ and $`\omega _k\omega _k^{}=dF_k`$, then $`\{u,v\}_\nu `$ and $`T^1\{Tu,Tv\}_\nu ^{}`$, with $`T=\mathrm{exp}DDu=\nu ^{k1}F_k(X_u)`$, correspond to forms which coincide to order $`k`$ so one has equivalence if all forms are cohomologous.
Reciprocally, if $`P^\mathrm{\Omega }`$ and $`P^\mathrm{\Omega }^{}`$ are equivalent and $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ coincide to order $`k1`$, then the equivalence can be written $`T=\mathrm{exp}DDu=_{jk}\nu ^jF_j(X_u)`$ and $`\omega _k\omega _k^{}=dF_k`$. Indeed if $`Tu=u+\nu ^rF_r(X_u)+\mathrm{}`$ , then the relation $`P^\mathrm{\Omega }(du,dv)=T^1P^\mathrm{\Omega }^{}(dTu,dTv)`$ at order $`rk`$ yields $`\omega _r\omega _r^{}=dF_r`$. When $`r<k`$ this means $`dF_r=0`$; locally, on a contractible set $`U`$, $`F_{r|U}=df_U`$. The map $`D_r:uP^\mathrm{\Omega }^{}(df_U,du)`$ is globally defined and is a derivation of $`\{,\}^{}`$ so $`T^{}=\mathrm{exp}\nu ^rD_rT=Id+\nu ^{r+1}F_{r+1}+\mathrm{}`$ is still an equivalence. By induction, one gets the result.
$``$In 1978, Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer stressed that different orderings in physics lead to equivalent star products on $`^{2n}`$. This shows that the notion of mathematical equivalence is not the same as the notion of physical equivalence (i.e. two star products leading to the same spectrum for each observable); we studied this difference with Cahen, Flato and Sternheimer in . Bayen et al. also proved in that Moyal star product is the only star product whose cochains are given by polynomials in the Poisson structure $`P`$; this was the first consideration of some universality property to build and classify star products.
$``$ Recall that a star product $``$ on $`(M,\omega )`$ is called differential if the $`2`$-cochains $`C_r(u,v)`$ giving it are bi-differential operators. As was observed by Lichnerowicz and Deligne :
###### Proposition 14
If $``$ and $`^{}`$ are differential star products and $`T(u)=u+_{r1}\nu ^rT_r(u)`$ is an equivalence so that $`T(uv)=T(u)^{}T(v)`$ then the $`T_r`$ are differential operators.
###### Proof.
Indeed if $`T=Id+\nu ^kT_k+\mathrm{}`$ then $`T_k=C_k^{}C_k`$ is differential so $`C_k^{}C_k`$ is a differential $`2`$-cocycle with vanishing skewsymmetric part but then, using Veyโs formula, it is the coboundary of a differential $`1`$-cochain $`E`$ and $`T_kE`$, being a $`1`$-cocycle, is a vector field so $`T_k`$ is differential. One then proceeds by induction, considering $`T^{}=(Id+\nu ^kT_k)^1T=Id+\nu ^{k+1}T_{k+1}^{}+\mathrm{}`$ and the two differential star products $``$ and $`^{\prime \prime }`$, where $`u^{\prime \prime }v=(Id+\nu ^kT_k)^1((Id+\nu ^kT_k)u^{}(Id+\nu ^kT_k)v),`$ which are equivalent through $`T^{}`$ (i.e. $`T^{}(uv)=T^{}(u)^{\prime \prime }T^{}(v)`$). โ
$``$ A differential star product is equivalent to one with linear term in $`\nu `$ given by $`\frac{1}{2}\{u,v\}`$. Indeed $`C_1(u,v)`$ is a Hochschild cocycle with antisymmetric part given by $`\frac{1}{2}\{u,v\}`$ so $`C_1=\frac{1}{2}P+B`$ for a differential $`1`$-cochain $`B`$. Setting $`T(u)=u+\nu B(u)`$ and $`u^{}v=T(T^1(u)T^1(v))`$, this equivalent star product $`^{}`$ has the required form.
$``$ In 1979, we proved that all differential deformed brackets on $`^{2n}`$ (or on any symplectic manifold such that $`b_2=0`$) are equivalent modulo a change of the parameter, and this implies a similar result for star products; this was proven by direct methods by Lichnerowicz :
###### Proposition 15
Let $``$ and $`^{}`$ be two differential star products on $`(M,\omega )`$ and suppose that $`H^2(M;)=0`$. Then there exists a local equivalence $`T=Id+_{k1}\nu ^kT_k`$ on $`C^{\mathrm{}}(M)[[\nu ]]`$ such that $`u^{}v=T(T^1uT^1v)`$ for all $`u,vC^{\mathrm{}}(M)[[\nu ]]`$.
###### Proof.
Let us suppose that, modulo some equivalence, the two star products $``$ and $`^{}`$ coincide up to order $`k`$. Then associativity at order $`k`$ shows that $`C_kC_k^{}`$ is a Hochschild $`2`$-cocycle and so by (4) can be written as $`(C_kC_k^{})(u,v)=(B)(u,v)+A(X_u,X_v)`$ for a $`2`$-form $`A`$. The total skewsymmetrization of the associativity relation at order $`k+1`$ shows that $`A`$ is a closed $`2`$-form. Since the second cohomology vanishes, $`A`$ is exact, $`A=dF`$. Transforming by the equivalence defined by $`Tu=u+\nu ^{k1}2F(X_u)`$, we can assume that the skewsymmetric part of $`C_kC_k^{}`$ vanishes. Then $`C_kC_k^{}=B`$ where $`B`$ is a differential operator. Using the equivalence defined by $`T=I+\nu ^kB`$ we can assume that the star products coincide, modulo an equivalence, up to order $`k+1`$ and the result follows from induction since two star products always agree in their leading term. โ
$``$ It followed from the above proof and results similar to (i.e. two star products which are equivalent and coincide at order $`k`$ differ at order $`k+1`$ by a Hochschild 2-cocycle whose skewsymmetric part corresponds to an exact 2-form) that at each step in $`\nu `$, equivalence classes of differential star products on a symplectic manifold $`(M,\omega )`$ are parametrised by $`H^2(M;)`$, if all such deformations exist. The general existence was proven by De Wilde and Lecomte. At that time, one assumed the parity condition $`C_n(u,v)=(1)^nC_n(v,u)`$, so equivalence classes of such differential star products were parametrised by series $`H^2(M;)[[\nu ^2]]`$. The parametrization was not canonical.
$``$ In 1994, Fedosov proved that his recursive construction works in a more general setting : given any series of closed 2-forms on a symplectic manifold $`(M,\omega )`$, he could build a connection on the Weyl bundle whose curvature is linked to that series and a star product whose equivalence class only depends on the element in $`H^2(M;)[[\nu ]]`$ corresponding to that series of forms.
$``$ In 1995, Nest and Tsygan , then Deligne and Bertelson proved that any differential star product on a symplectic manifold $`(M,\omega )`$ is equivalent to a Fedosov star product and that its equivalence class is parametrised by the corresponding element in $`H^2(M;)[[\nu ]]`$.
$``$ In 1997, Kontsevich proved that the coincidence of the set of equivalence classes of star and Poisson deformations is true for general Poisson manifolds :
###### Theorem 16
The set of equivalence classes of differential star products on a Poisson manifold $`(M,P)`$ can be naturally identified with the set of equivalence classes of Poisson deformations of $`P`$:
$$P_\nu =P\nu +P_2\nu ^2+\mathrm{}\mathrm{\Gamma }(X,^2T_X)[[\nu ]],[P_\nu ,P_\nu ]=0.$$
$``$ Remark that all results concerning parametrisation of equivalence classes of differential star products are still valid for star products defined by local cochains or for star products defined by continuous cochains (, Pinczon ). Parametrization of equivalence classes of special star products have been obtained : star products with separation of variables (by Karabegov ), invariant star products on a symplectic manifold when there exists an invariant symplectic connection (with Bertelson and Bieliavsky ), algebraic star products (Chloup , Kontsevich )โฆ
$``$ The association of an element in $`H^2(M;)[[\nu ]]`$ to the equivalence class of a star product on a symplectic manifold is one way to associate an invariant to a star product; other such associations are obtained by star version of index theorems and trace functionals on the algebra $`(C^{\mathrm{}}(M)[[\nu ]],)`$. I shall not develop that aspect here. It was first considered by Connes, Flato and Sternheimer in where they introduce the notion of closed star product, i.e. such that
$$_Mab\omega ^n=_Mba\omega ^n(\mathrm{mod}\nu ^n)a,bC^{\mathrm{}}(M)[[\nu ]],$$
and show how their classification is linked to cyclic cohomology.
They obtain, for the cotangent bundle to a compact Riemannian manifold, its Todd class as the โcharacterโ associated to the star product corresponding to normal ordering.
The notion of a trace for star products and the star version of index theorems have been studied by Fedosov and by Nest and Tsygan .
### 3.1 Deligneโs cohomology classes associated to differential star products on symplectic manifolds
Deligne defines two cohomological classes associated to differential star products on a symplectic manifold. This leads to an intrinsic way to parametrise the equivalence class of such a differential star product. Although the question makes sense more generally for Poisson manifolds, Deligneโs method depends crucially on the Darboux theorem and the uniqueness of the Moyal star product on $`^{2n}`$ so the methods do not extend to general Poisson manifolds.
The first class is a relative class; fixing a star product on the manifold, it intrinsically associates to any equivalence class of star products an element in $`H^2(M;)[[\nu ]]`$. This is done in ฤech cohomology by looking at the obstruction to gluing local equivalences.
Deligneโs second class is built from special local derivations of a star product. The same derivations played a special role in the first general existence theorem for a star product on a symplectic manifold. Deligne used some properties of Fedosovโs construction and central curvature class to relate his two classes and to see how to characterise an equivalence class of star products by the derivation related class and some extra data obtained from the second term in the deformation. With John Rawnsley , we did this by direct ฤech methods which I shall present here.
#### 3.1.1 The relative class
Let $``$ and $`^{}`$ be two differential star products on $`(M,\omega )`$. Let $`U`$ be a contractible open subset of $`M`$ and $`N_U=C^{\mathrm{}}(U)`$. Remark that any differential star product on $`M`$ restricts to $`U`$ and $`H^2(M;)(U)=0`$, hence, by proposition 15, there exists a local equivalence $`T=Id+_{k1}\nu ^kT_k`$ on $`N_U[[\nu ]]`$ so that $`u^{}v=T(T^1uT^1v)`$ for all $`u,vN_U[[\nu ]]`$.
###### Proposition 17
Let $``$ be a differential star product on $`(M,\omega )`$ and suppose that $`H^1(M;)`$ vanishes.
* Any self-equivalence $`A=Id+_{k1}\nu ^kA_k`$ of $``$ is inner: $`A=\mathrm{exp}ad_{}a`$ for some $`aC^{\mathrm{}}(M)[[\nu ]]`$.
* Any $`\nu `$-linear derivation of $``$ is of the form $`D=_{i0}\nu ^iD_i`$ where each $`D_i`$ corresponds to a symplectic vector field $`X_i`$ and is given on a contractible open set $`U`$ by
$$D_iu|_U=\frac{1}{\nu }\left(f_i^Uuuf_i^U\right)$$
if $`X_iu|_U=\{f_i^U,u\}|_U`$.
Indeed, one builds $`a`$ recursively; assuming $`A=Id+_{rk}\nu ^rA_r`$ and $`k1`$, the condition $`A(uv)=AuAv`$ implies at order $`k`$ in $`\nu `$ that $`A_k(uv)+C_k(u,v)=A_k(u)v+uA_k(v)+C_k(u,v)`$ so that $`A_k`$ is a vector field. Taking the skew part of the terms in $`\nu ^{k+1}`$ we have that $`A_k`$ is a derivation of the Poisson bracket. Since $`H^1(M;)=0`$, one can write $`A_k(u)=\{a_{k1},u\}`$ for some function $`a_{k1}`$. Then $`(\mathrm{exp}ad_{}\nu ^{k1}a_{k1})A=Id+O(\nu ^{k+1})`$ and the induction proceeds. The proof for $`\nu `$-linear derivation is similar.
The above results can be applied to the restriction of a differential star product on $`(M,\omega )`$ to a contractible open set $`U`$. Set, as above, $`N_U=C^{\mathrm{}}(U)`$. If $`A=Id+_{k1}\nu ^kA_k`$ is a formal linear operator on $`N_U[[\nu ]]`$ which preserves the differential star product $``$, then there is $`aN_U[[\nu ]]`$ with $`A=\mathrm{exp}ad_{}a`$. Similarly, any local $`\nu `$-linear derivation $`D_U`$ of $``$ on $`N_U[[\nu ]]`$ is essentially inner: $`D_U=\frac{1}{\nu }ad_{}d_U`$ for some $`d_UN_U[[\nu ]]`$.
It is convenient to write the composition of automorphisms of the form $`\mathrm{exp}ad_{}a`$ in terms of $`a`$. In a pronilpotent situation this is done with the CampbellโBakerโHausdorff composition which is denoted by $`a_{}b`$:
$$a_{}b=a+_0^1\psi (\mathrm{exp}ad_{}a\mathrm{exp}tad_{}b)b๐t$$
where
$$\psi (z)=\frac{z\mathrm{log}(z)}{z1}=\underset{n1}{}\left(\frac{(1)^n}{n+1}+\frac{(1)^{n+1}}{n}\right)(z1)^n.$$
Notice that the formula is well defined (at any given order in $`\nu `$, only a finite number of terms arise) and it is given by the usual series
$$a_{}b=a+b+\frac{1}{2}[a,b]_{}+\frac{1}{12}([a,[a,b]_{}]_{}+[b,[b,a]_{}]_{})\mathrm{}.$$
The following results are standard (N. Bourbaki, Groupes et algรจbres de Lie, รlรฉments de Mathรฉmatique, Livre 9, Chapitre 2, ยง6):
* $`_{}`$ is an associative composition law;
* $`\mathrm{exp}ad_{}(a_{}b)=\mathrm{exp}ad_{}a\mathrm{exp}ad_{}b`$;
* $`a_{}b_{}(a)=\mathrm{exp}(ad_{}a)b`$;
* $`(a_{}b)=(b)_{}(a)`$;
* $`{\displaystyle \frac{d}{dt}}|_0(a)_{}(a+tb)={\displaystyle \frac{1\mathrm{exp}(ad_{}a)}{ad_{}a}}(b)`$.
Let $`(M,\omega )`$ be a symplectic manifold. We fix a locally finite open cover $`๐ฐ=\{U_\alpha \}_{\alpha I}`$ by Darboux coordinate charts such that the $`U_\alpha `$ and all their non-empty intersections are contractible, and we fix a partition of unity $`\{\theta _\alpha \}_{\alpha I}`$ subordinate to $`๐ฐ`$. Set $`N_\alpha =C^{\mathrm{}}(U_\alpha )`$, $`N_{\alpha \beta }=C^{\mathrm{}}(U_\alpha U_\beta )`$, and so on.
Now suppose that $``$ and $`^{}`$ are two differential star products on $`(M,\omega )`$. We have seen that their restrictions to $`N_\alpha [[\nu ]]`$ are equivalent so there exist formal differential operators $`T_\alpha :N_\alpha [[\nu ]]N_\alpha [[\nu ]]`$ such that
$$T_\alpha (uv)=T_\alpha (u)^{}T_\alpha (v),u,vN_\alpha [[\nu ]].$$
On $`U_\alpha U_\beta `$, $`T_\beta ^1T_\alpha `$ will be a self-equivalence of $``$ on $`N_{\alpha \beta }[[\nu ]]`$ and so there will be elements $`t_{\beta \alpha }=t_{\alpha \beta }`$ in $`N_{\alpha \beta }[[\nu ]]`$ with
$$T_\beta ^1T_\alpha =\mathrm{exp}ad_{}t_{\beta \alpha }.$$
On $`U_\alpha U_\beta U_\gamma `$ the element
$$t_{\gamma \beta \alpha }=t_{\alpha \gamma }_{}t_{\gamma \beta }_{}t_{\beta \alpha }$$
induces the identity automorphism and hence is in the centre $`[[\nu ]]`$ of $`N_{\alpha \beta \gamma }[[\nu ]]`$. The family of $`t_{\gamma \beta \alpha }`$ is thus a ฤech $`2`$-cocycle for the covering $`๐ฐ`$ with values in $`[[\nu ]]`$. The standard arguments show that its class does not depend on the choices made, and is compatible with refinements. Since every open cover has a refinement of the kind considered it follows that $`t_{\gamma \beta \alpha }`$ determines a unique ฤech cohomology class $`[t_{\gamma \beta \alpha }]H^2(M;)[[\nu ]]`$.
###### Definition 18
$$t(^{},)=[t_{\gamma \beta \alpha }]H^2(M;)[[\nu ]]$$
is Deligneโs relative class.
It is easy to see, using the fact that the cohomology of the sheaf of smooth functions is trivial:
###### Theorem 19
(Deligne) Fixing a differential star product $``$, the class $`t(^{},)`$ in $`H^2(M;)[[\nu ]]`$ depends only on the equivalence class of the differential star product $`^{}`$, and sets up a bijection between the set of equivalence classes of differential star products and $`H^2(M;)[[\nu ]]`$.
If $``$, $`^{}`$, $`^{\prime \prime }`$ are three differential star products on $`(M,\omega )`$ then
$$t(^{\prime \prime },)=t(^{\prime \prime },^{})+t(^{},).$$
(4)
#### 3.1.2 The derivation related class
The addition formula above suggests that $`t(^{},)`$ should be a difference of classes $`c(^{}),c()H^2(M;)[[\nu ]]`$. Moreover, the class $`c()`$ should determine the star product $``$ up to equivalence.
###### Definition 20
Let $`U`$ be an open set of $`M`$. Say that a derivation $`D`$ of $`(C^{\mathrm{}}(U)[[\nu ]],)`$ is $`\nu `$-Euler if it has the form
$$D=\nu \frac{}{\nu }+X+D^{}$$
(5)
where $`X`$ is conformally symplectic on $`U`$ ($`_X\omega |_U=\omega |_U`$) and $`D^{}=_{r1}\nu ^rD_r^{}`$ with the $`D_r^{}`$ differential operators on $`U`$.
###### Proposition 21
Let $``$ be a differential star product on $`(M,\omega )`$. For each $`U_\alpha ๐ฐ`$ there exists a $`\nu `$-Euler derivation $`D_\alpha =\nu \frac{}{\nu }+X_\alpha +D_\alpha ^{}`$ of the algebra $`(N_\alpha [[\nu ]],)`$.
###### Proof.
On an open set in $`^{2n}`$ with the standard symplectic structure $`\mathrm{\Omega }`$, denote the Poisson bracket by $`P`$. Let $`X`$ be a conformal vector field so $`_X\mathrm{\Omega }=\mathrm{\Omega }`$. The Moyal star product $`_M`$ is given by $`u_Mv=uv+_{r1}(\frac{\nu }{2})^r/r!P^r(u,v)`$ and $`D=\nu \frac{}{\nu }+X`$ is a derivation of $`_M`$.
Now $`(U_\alpha ,\omega )`$ is symplectomorphic to an open set in $`^{2n}`$ and any differential star product on this open set is equivalent to $`_M`$ so we can pull back $`D`$ and $`_M`$ to $`U_\alpha `$ by a symplectomorphism to give a star product $`^{}`$ with a derivation of the form $`\nu \frac{}{\nu }+X_\alpha `$. If $`T`$ is an equivalence of $``$ with $`^{}`$ on $`U_\alpha `$ then $`D_\alpha =T^1(\nu \frac{}{\nu }+X_\alpha )T`$ is a derivation of the required form. โ
We take such a collection of derivations $`D_\alpha `$ given by Proposition 21 and on $`U_\alpha U_\beta `$ we consider the differences $`D_\beta D_\alpha `$. They are derivations of $``$ and the $`\nu `$ derivatives cancel out, so $`D_\beta D_\alpha `$ is a $`\nu `$-linear derivation of $`N_{\alpha \beta }[[\nu ]]`$. Any $`\nu `$-linear derivation is of the form $`\frac{1}{\nu }ad_{}d`$, so there are $`d_{\beta \alpha }N_{\alpha \beta }[[\nu ]]`$ with
$$D_\beta D_\alpha =\frac{1}{\nu }ad_{}d_{\beta \alpha }$$
(6)
with $`d_{\beta \alpha }`$ unique up to a central element. On $`U_\alpha U_\beta U_\gamma `$ the combination $`d_{\alpha \gamma }+d_{\gamma \beta }+d_{\beta \alpha }`$ must be central and hence defines $`d_{\gamma \beta \alpha }[[\nu ]]`$. It is easy to see that $`d_{\gamma \beta \alpha }`$ is a $`2`$-cocycle whose ฤech class $`[d_{\gamma \beta \alpha }]H^2(M;)[[\nu ]]`$ does not depend on any of the choices made.
###### Definition 22
$`d()=[d_{\gamma \beta \alpha }]H^2(M;)[[\nu ]]`$ is Deligneโs intrinsic derivation-related class.
* In fact the class considered by Deligne is actually $`\frac{1}{\nu }d()`$. A purely ฤech-theoretic accounts of this class is given in Karabegov .
* If $``$ and $`^{}`$ are equivalent then $`d(^{})=d()`$.
* If $`d()=_{r0}\nu ^rd^r()`$ then $`d^0()=[\omega ]`$ under the de Rham isomorphism, and $`d^1()=0`$.
Consider two differential star products $``$ and $`^{}`$ on $`(M,\omega )`$ with local equivalences $`T_\alpha `$ and local $`\nu `$-Euler derivations $`D_\alpha `$ for $``$ . Then $`D_\alpha ^{}=T_\alpha D_\alpha T_\alpha ^1`$ are local $`\nu `$-Euler derivations for $`^{}`$. Let $`D_\beta D_\alpha =\frac{1}{\nu }ad_{}d_{\beta \alpha }`$ and $`T_\beta ^1T_\alpha =\mathrm{exp}ad_{}t_{\beta \alpha }`$ on $`U_\alpha U_\beta `$. Then $`D_\beta ^{}D_\alpha ^{}=\frac{1}{\nu }ad_{^{}}d_{\beta \alpha }^{}`$ where
$$d_{\beta \alpha }^{}=T_\beta d_{\beta \alpha }\nu T_\beta \left(\frac{1\mathrm{exp}(ad_{}t_{\alpha \beta })}{ad_{}t_{\alpha \beta }}\right)D_\alpha t_{\alpha \beta }.$$
In this situation
$$d_{\gamma \beta \alpha }^{}=T_\alpha (d_{\gamma \beta \alpha }+\nu ^2\frac{}{\nu }t_{\gamma \beta \alpha }).$$
This gives a direct proof of:
###### Theorem 23
(Deligne) The relative class and the intrinsic derivation-related classes of two differential star products $``$ and $`^{}`$ are related by
$$\nu ^2\frac{}{\nu }t(^{},)=d(^{})d().$$
(7)
#### 3.1.3 The characteristic class
The formula above shows that the information which is โlostโ in $`d(^{})d()`$ corresponds to the zeroth order term in $`\nu `$ of $`t(^{},)`$.
###### Remark 24
In it was shown that any bidifferential operator $`C`$, vanishing on constants, which is a $`2`$-cocycle for the Chevalley cohomology of $`(C^{\mathrm{}}(M),\{,\})`$ with values in $`C^{\mathrm{}}(M)`$ associated to the adjoint representation (i.e. such that $`S_{u,v,w}[\{u,C(v,w)\}C(\{u,v\},w)]=0`$ where $`S_{u,v,w}`$ denotes the sum over cyclic permutations of $`u,v`$ and $`w`$) can be written as
$$C(u,v)=aS_\mathrm{\Gamma }^3(u,v)+A(X_u,X_v)+[\{u,Ev\}+\{Eu,v\}E(\{u,v\})]$$
where $`a`$, where $`S_\mathrm{\Gamma }^3`$ is a bidifferential $`2`$-cocycle introduced in (which vanishes on constants and is never a coboundary and whose symbol is of order $`3`$ in each argument), where $`A`$ is a closed $`2`$-form on $`M`$ and where $`E`$ is a differential operator vanishing on constants. Hence
$$H_{\mathrm{Chev},\mathrm{nc}}^2(C^{\mathrm{}}(M),C^{\mathrm{}}(M))=H^2(M;)$$
and we define the $`\mathrm{\#}`$ operator as the projection on the second factor relative to this decomposition.
###### Proposition 25
Given two differential star products $``$ and $`^{}`$, the term of order zero in Deligneโs relative class $`t(^{},)=_{r0}\nu ^rt^r(^{},)`$ is given by
$$t^0(^{},)=2(C_{}^{}{}_{2}{}^{})^\mathrm{\#}+2(C_2^{})^\mathrm{\#}.$$
If $`C_1=\frac{1}{2}\{,\}`$, then $`C_2^{}(u,v)=A(X_u,X_v)`$ where $`A`$ is a closed $`2`$-form and $`(C_2^{})^\mathrm{\#}=[A]`$ so it โisโ the skewsymmetric part of $`C_2`$.
It follows from what we did before that the association to a differential star product of $`(C_2^{})^\mathrm{\#}`$ and $`d()`$ completely determines its equivalence class.
###### Definition 26
The characteristic class of a differential star product $``$ on $`(M,\omega )`$ is the element $`c()`$ of the affine space $`\frac{[\omega ]}{\nu }+H^2(M;)[[\nu ]]`$ defined by
$`c()^0`$ $`=`$ $`2(C_2^{})^\mathrm{\#}`$
$`{\displaystyle \frac{}{\nu }}c()(\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\nu ^2}}d()`$
###### Theorem 27
The characteristic class has the following properties:
* The relative class is given by
$$t(^{},)=c(^{})c()$$
(8)
* The map $`C`$ from equivalence classes of star products on $`(M,\omega )`$ to the affine space $`\frac{[\omega ]}{\nu }+H^2(M;)[[\nu ]]`$ mapping $`[]`$ to $`c()`$ is a bijection.
* If $`\psi :MM^{}`$ is a diffeomorphism and if $``$ is a star product on $`(M,\omega )`$ then $`u^{}v=(\psi ^1)^{}(\psi ^{}u\psi ^{}v)`$ defines a star product denoted $`^{}=(\psi ^1)^{}`$ on $`(M^{},\omega ^{})`$ where $`\omega ^{}=(\psi ^1)^{}\omega `$. The characteristic class is natural relative to diffeomorphisms:
$$c((\psi ^1)^{})=(\psi ^1)^{}c().$$
(9)
* Consider a change of parameter $`f(\nu )=_{r1}\nu ^rf_r`$ where $`f_r`$ and $`f_10`$ and let $`^{}`$ be the star product obtained from $``$ by this change of parameter, i.e. $`u^{}v=u.v+_{r1}(f(\nu ))^rC_r(u,v)=u.v+f_1\nu C_1(u,v)+\nu ^2((f_1)^2C_2(u,v)+f_2C_1(u,v))+\mathrm{}`$. Then $`^{}`$ is a differential star product on $`(M,\omega ^{})`$ where $`\omega ^{}=\frac{1}{f_1}\omega `$ and we have equivariance under a change of parameter:
$$c(^{})(\nu )=c()(f(\nu )).$$
(10)
The characteristic class $`c()`$ coincides (cf Deligne and Neumaier ) for Fedosov-type star products with their characteristic class introduced by Fedosov as the de Rham class of the curvature of the generalised connection used to build them (up to a sign and factors of $`2`$). That characteristic class is also studied by Weinstein and Xu in . The fact that $`d()`$ and $`(C_2^{})^\mathrm{\#}`$ completely characterise the equivalence class of a star product is also proven by ฤech methods in De Wilde .
### 3.2 Automorphisms of a star product and generalised moment map
The above proposition allows to study automorphisms of star products on a symplectic manifold (, ).
###### Definition 28
An isomorphism from a differential star product $``$ on $`(M,\omega )`$ to a differential star product $`^{}`$ on $`(M^{},\omega ^{})`$ is an $``$-linear bijective map $`A:C^{\mathrm{}}(M)[[\nu ]]C^{\mathrm{}}(M^{})[[\nu ]]`$, continuous in the $`\nu `$-adic topology (i.e. $`A(_r\nu ^ru_r)`$ is the limit of $`_{rN}A(\nu ^ru_r)`$ ), such that
$$A(uv)=Au^{}Av.$$
Notice that if $`A`$ is such an isomorphism, then $`A(\nu )`$ is central for $`^{}`$ so that $`A(\nu )=f(\nu )`$ where $`f(\nu )[[\nu ]]`$ is without constant term to get the $`\nu `$-adic continuity. Let us denote by $`^{\prime \prime }`$ the differential star product on $`(M,\omega _1=\frac{1}{f_1}\omega )`$ obtained by a change of parameter
$$u_\nu ^{^{\prime \prime }}v=u_{f(\nu )}v=F(F^1uF^1v)$$
for $`F:C^{\mathrm{}}(M)[[\nu ]]C^{\mathrm{}}(M)[[\nu ]]:_r\nu ^ru_r_rf(\nu )^ru_r`$.
Define $`A^{}:C^{\mathrm{}}(M)[[\nu ]]C^{\mathrm{}}(M^{})[[\nu ]]`$ by $`A=A^{}F`$. Then $`A^{}`$ is a $`\nu `$-linear isomorphism between $`^{\prime \prime }`$ and $`^{}`$:
$$A^{}(u^{\prime \prime }v)=A^{}u^{}A^{}v.$$
At order zero in $`\nu `$ this yields $`A_0^{}(u.v)=A_0^{}u.A_0^{}v`$ so that there exists a diffeomorphism $`\psi :M^{}M`$ with $`A_0^{}u=\psi ^{}u`$. The skewsymmetric part of the isomorphism relation at order $`1`$ in $`\nu `$ implies that $`\psi ^{}\omega _1=\omega ^{}`$. Let us denote by $`^{\prime \prime \prime }`$ the differential star product on $`(M,\omega _1)`$ obtained by pullback via $`\psi `$ of $`^{}`$:
$$u^{\prime \prime \prime }v=(\psi ^1)^{}(\psi ^{}u^{}\psi ^{}v)$$
and define $`B:C^{\mathrm{}}(M)[[\nu ]]C^{\mathrm{}}(M)[[\nu ]]`$ so that $`A^{}=\psi ^{}B`$. Then $`B`$ is $`\nu `$-linear, starts with the identity and
$$B(u^{\prime \prime }v)=Bu^{\prime \prime \prime }Bv$$
so that $`B`$ is an equivalence โ in the usual sense โ between $`^{\prime \prime }`$ and $`^{\prime \prime \prime }`$. Hence
###### Proposition 29
Any isomorphism between two differential star products on symplectic manifolds is the combination of a change of parameter and a $`\nu `$-linear isomorphism. Any $`\nu `$-linear isomorphism between two star products $``$ on $`(M,\omega )`$ and $`^{}`$ on $`(M^{},\omega ^{})`$ is the combination of the action on functions of a symplectomorphism $`\psi :M^{}M`$ and an equivalence between $``$ and the pullback via $`\psi `$ of $`^{}`$. In particular, it exists if and only if those two star products are equivalent, i.e. if and only if $`(\psi ^1)^{}c(^{})=c()`$, where here $`(\psi ^1)^{}`$ denotes the action on the second de Rham cohomology space.
In particular, two differential star products $``$ on $`(M,\omega )`$ and $`^{}`$ on $`(M^{},\omega ^{})`$ are isomorphic if and only if there exist $`f(\nu )=_{r1}\nu ^rf_r[[\nu ]]`$ with $`f_10`$ and $`\psi :M^{}M`$, a symplectomorphism, such that $`(\psi ^1)^{}c(^{})(f(\nu ))=c()(\nu )`$. In particular : if $`H^2(M;)=[\omega ]`$ then there is only one star product up to equivalence and change of parameter.
Omori et al. also show that when reparametrizations are allowed then there is only one star product on $`P^n`$.
A special case of Proposition 29 gives:
###### Proposition 30
A symplectomorphism $`\psi `$ of a symplectic manifold can be extended to a $`\nu `$-linear automorphism of a given differential star product on $`(M,\omega )`$ if and only if $`(\psi )^{}c()=c()`$.
Notice that this is always the case if $`\psi `$ can be connected to the identity by a path of symplectomorphisms (and this result was in Fedosov ).
$``$ If $`G`$ is a connected Lie group acting on the symplectic manifold $`(M,\omega )`$ by symplectomorphisms, each element of $`G`$ can be lifted to an automorphism of a star product on $`M`$. The group $`G`$ acts on the quantum level if there is a homomorphism $`\rho `$ from $`G`$ into the automorphism group of $``$ such that $`\rho (g)=g^{}+\nu \mathrm{}gG`$. At the Lie algebra level, one considers a homomorphism $`\sigma `$ from the Lie algebra $`๐ค`$ of $`G`$ to the algebra of derivations of $``$ such that $`\sigma (X)=X^{}+\nu \mathrm{}X๐ค`$ where $`X^{}`$ is the fundamental vector field on $`M`$ associated to the action of $`G`$ on $`M`$ (i.e. $`X_x^{}=\frac{d}{dt}\mathrm{exp}tXx_{|_0}).`$ Now any local $`\nu `$-linear derivation of $``$ can be locally written on a contractible set $`U`$ as $`ad_{}(\frac{1}{\nu }f)`$ for some $`fC^{\mathrm{}}(U)[[\nu ]]`$.
###### Definition 31
Given a star product $``$ on a Poisson manifold $`(M,P)`$ and given a connected Lie group $`G`$ acting on $`M`$, the star product gives a quantization of the action of the Lie algebra $`๐ค`$ if there exists a generalised moment map i.e. a map
$$\tau :๐ค\frac{1}{\nu }C^{\mathrm{}}(M)[[\nu ]]:X\tau (X)$$
such that
$$\tau (X)\tau (Y)\tau (Y)\tau (X)=\tau ([X,Y])X,Y๐ค$$
(11)
and
$$\tau (X)uu\tau (X)=X^{}u+\nu \mathrm{}X๐ค.$$
(12)
Remark that the two conditions imply for the term of $`\tau `$ of order $`1`$ in $`\nu `$ that the action of $`G`$ on $`M`$ admits a moment map, i.e. there is a map $`\lambda :๐คC^{\mathrm{}}(M)`$ such that $`\{\lambda (Y),u\}=Y^{}u`$ and $`\{\lambda (Y),\lambda (Y)\}=\lambda (\{X,Y\})`$, and
$$\tau (X)=\frac{1}{\nu }\lambda (X)+\mathrm{}.$$
When one can choose $`\tau (X)=\frac{1}{\nu }\lambda (X)X๐ค`$ the notion of a star product which is a quantization of the action of $`๐ค`$ coincides with the notion of a covariant star product.
When one can choose $`\tau `$ so that $`\tau (X)uu\tau (X)=X^{}uX๐ค`$, it implies that $`X^{}(uv)=(X^{}u)v+u(X^{}v)`$, so that
$$g^{}(uv)=g^{}ug^{}vgG$$
which are the conditions in Bayen et al for the star product to be geometrically invariant; in that situation, the notion of generalised moment map coincides with the notion of quantum moment map introduced by Xu in .
In the general case, when one has found a map $`\tau `$ at order $`k`$ in $`\nu `$ satisfying condition (11) at order $`k`$, then one can extend things one order further if a Chevalley $`2`$-cocycle from $`๐ค`$ with values in $`C^{\mathrm{}}(M)`$ (for the representation of $`Y๐ค`$ given by $`Y^{}`$) is a $`2`$-coboundary. When one deals with geometrically invariant star products, one can always find a $`\tau `$ such that $`\tau (X)uu\tau (X)=X^{}uX๐ค`$ if $`H^2(๐ค,)=0`$ and $`H^1(M;)=0`$ as was found in .
## 4 Convergence of star products
###### Remark 32
Let $`(M,P)`$ be a Poisson manifold and let $``$ be a differential star product on it with $`1`$ acting as the identity. Observe that if there exists a value $`k`$ of $`\nu `$ such that
$$uv=\underset{r=0}{\overset{\mathrm{}}{}}\nu ^rC_r(u,v)$$
converges (for the pointwize convergence of functions), for all $`u,vC^{\mathrm{}}(M)`$, to $`F_k(u,v)`$ in such a way that $`F_k`$ is associative, then $`F_k(u,v)=uv`$.
So assuming โtoo muchโ convergence kills all deformations. On the other hand, in any physical situation, one needs some convergence properties to be able to compute the spectrum of quantum observables in terms of a star product as was done for some observables already in Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer .
Now consider the example of Moyal star product on the symplectic vector space $`(^{2n},\omega )`$. The formal formula for Moyal star product
$$(u_Mv)(z)=\mathrm{exp}\left(\frac{\nu }{2}P^{rs}_{x^r}_{y^s}\right)(u(x)v(y))|_{x=y=z}$$
obviously converges when $`u`$ and $`v`$ are polynomials.
On the other hand, there is an integral formula for Moyal star product given by
$$(uv)(\xi )=(\pi \mathrm{})^{2n}u(\xi ^{})v(\xi ^{\prime \prime })e^{\frac{2i}{\mathrm{}}(\omega (\xi ,\xi ^{\prime \prime })+\omega (\xi ^{\prime \prime },\xi ^{})+\omega (\xi ^{},\xi ))}๐\xi ^{}๐\xi ^{\prime \prime },$$
and this product $``$ gives a structure of associative algebra on the space of rapidly decreasing functions $`๐ฎ(^{2n})`$.
The formal formula converges (for $`\nu =i\mathrm{}`$) in the topology of $`๐ฎ^{}`$ for $`u`$ and $`v`$ with compact Fourier transform.
With Michel Cahen and John Rawnsley, we used the method of quantization of Kรคhler manifolds due to Berezin, , as the inverse of taking symbols of operators, to construct on Hermitian symmetric spaces star products which are convergent on a large class of functions on the manifold. I shall develop this construction in this last part of my talk.
Let me mention, before closing this introduction about convergence, the work of Rieffel where he introduces the notion of strict deformation quantization. An example of strict Frรฉchet quantization has been recently given by Omori, Maeda, Niyazaki and Yoshioka in .
Also very important are the constructions of operator representations of star products, in particular the works of Fedosov and of Bordemann, Neumaier and Waldmann .
### 4.1 Convergence of Berezin type star products on Hermitian symmetric spaces
The method to construct a star product involves making a correspondence between operators and functions (their Berezin symbols), transferring the operator composition to the symbols, introducing a suitable parameter into the Berezin composition of symbols, taking the asymptotic expansion in this parameter on a large algebra of functions and then showing that the coefficients of this expansion satisfy the cocycle conditions to define a star product on the smooth functions. The idea of an asymptotic expansion appeared already in Berezin and in Moreno and Ortega-Navarro.
In we show that this asymptotic expansion exists for compact $`M`$, and defines an associative multiplication on formal power series in $`k^1`$ with coefficients in $`C^{\mathrm{}}(M)`$ for compact coadjoint orbits. We also show that this formal power series converges on the space of symbols for $`M`$ a Hermitian symmetric space of compact type.
In , Karabegov proves convergence for general compact coadjoint orbits (i.e. flag manifolds).
In we study general Hermitian symmetric spaces of non-compact type, and use their realisation as bounded domains to define an analogous algebra of symbols of polynomial differential operators.
Recently Reshetikhin and Takhtajan have announced an associative formal star product given by an asymptotic expansion on any Kรคhler manifold. This they do in two steps, first building an associative product for which $`1`$ is not a unit element, then passing to a star product.
#### 4.1.1 Berezin symbols
We denote by $`(L,,h)`$ a quantization bundle for the Kรคhler manifold $`(M,\omega ,J)`$ (i.e. a holomorphic line bundle $`L`$ with connection $``$ admitting an invariant hermitian structure $`h`$, such that the curvature is curv$`()=2i\pi \omega `$). We denote by $``$ the Hilbert space of square-integrable holomorphic sections of $`L`$ which we assume to be non-trivial. The coherent states are vectors $`e_q`$ such that
$$s(x)=s,e_qq,q_x,xM,s$$
where $``$ denotes the complement of the zero-section in $`L`$. The function
$$ฯต(x)=|q|^2e_q^2,q_x$$
is well-defined and real analytic.
We introduce also the $`2`$-point function
$$\psi (x,y)=\frac{|e_q^{},e_q|^2}{e_q^{}^2e_q^2},q_x,q^{}_y$$
which is a globally defined real analytic function on $`M\times M`$ provided $`ฯต`$ has no zeros. It is a consequence of the CauchyโSchwartz inequality that $`\psi (x,y)1`$ everywhere, with equality where the lines spanned by $`e_q`$ and $`e_q^{}`$ coincide ($`q_x`$, $`q^{}_y`$).
Let $`A:`$ be a bounded linear operator and let
$$\widehat{A}(x)=\frac{Ae_q,e_q}{e_q,e_q},q_x,xM$$
be its symbol. The function $`\widehat{A}`$ has an analytic continuation to an open neighbourhood of the diagonal in $`M\times \overline{M}`$ given by
$$\widehat{A}(x,y)=\frac{Ae_q^{},e_q}{e_q^{},e_q},q_x,q^{}_y$$
which is holomorphic in $`x`$ and antiholomorphic in $`y`$. We denote by $`\widehat{E}(L)`$ the space of symbols of bounded operators on $``$. We can extend this definition of symbols to some unbounded operators provided everything is well defined.
#### 4.1.2 Composition of operators - Parameter
The composition of operators on $``$ gives rise to a product for the corresponding symbols, which is associative and which we shall denote by $``$ following Berezin, . The product $``$ of symbols is given in terms of the symbols by the integral formula
$$(\widehat{A}\widehat{B})(x)=_M\widehat{A}(x,y)\widehat{B}(y,x)\psi (x,y)ฯต(y)\frac{\omega ^n(y)}{n!}.$$
This formula is derived by use of the adjoint $`A^{}`$ of $`A`$ so to apply it to the case where the operators are unbounded we need to be able to use the adjoint of $`A`$ on coherent states. To be able to take the symbol of the composition, the result of applying $`B`$ to a coherent state must be in the domain of $`A`$.
Example: The identity map has symbol $`\widehat{I}=1`$ and so $`\widehat{A}1=1\widehat{A}=\widehat{A}`$ for any operator $`A`$. In particular $`11=1`$.
Let $`k`$ be a positive integer. The bundle $`(L^k=^kL,^k,h^k)`$ is a quantization bundle for $`(M,k\omega ,J)`$ and we denote by $`^k`$ the corresponding space of holomorphic sections and by $`\widehat{E}(L^k)`$ the space of symbols of linear operators on $`^k`$. We let $`ฯต^{(k)}`$ be the corresponding function. We say that the quantization is regular if $`ฯต^{(k)}`$ is a non-zero constant for all nonnegative $`k`$ and if $`\psi (x,y)=1`$ implies $`x=y`$. The significance of these conditions has been explained in .
Let $`G`$ be a Lie group of isometries of the Kรคhler manifold $`(M,\omega ,J)`$ which lifts to a group of automorphisms of the quantization bundle $`(L,,h)`$. This automorphism group acts naturally on $`(L^k,^{(k)},h^k)`$; if $`gG`$ and if $`e_q^{(l)}`$ is a coherent state of $`L^l`$, then $`g.e_q^{(l)}=e_{gq}^{(l)}`$ so the function $`ฯต^{(k)}`$ is invariant under $`G`$. In particular, if the quantization is homogeneous, all $`ฯต^{(k)}`$ are constants.
In the regular case, the function $`\psi `$ in the integral defining the composition of symbols for powers $`L^k`$ gets replaced by powers $`\psi ^k`$.
When the manifold is compact, we have proven the following facts:
\- when $`ฯต^{(k)}`$ is constant for all $`k`$ one has the nesting property $`\widehat{E}(L^k)\widehat{E}(L^{k+1})`$;
\- with the same assumption $`_k\widehat{E}(L^k)`$ is dense in $`๐^{}(M)`$.
From the nesting property, one sees that if $`\widehat{A},\widehat{B}`$ belong to $`\widehat{E}(L^l)`$ and if $`kl`$ one may define
$$(\widehat{A}_k\widehat{B})(x)=_M\widehat{A}(x,y)\widehat{B}(y,x)\psi ^k(x,y)ฯต^{(k)}\frac{k^n\omega ^n}{n!}.$$
(13)
More generally, if $`\widehat{A}`$ is a symbol of an operator then its analytic continuation $`\widehat{A}(x,y)`$ may have singularities where $`\psi (x,y)=0`$ but $`\widehat{A}\psi `$ is always globally defined on $`M\times M`$. If $`M`$ is not compact $`\widehat{A}\psi `$ may not be bounded, so we introduce the class $`C^{\mathrm{}}(M)`$ of functions $`f`$ which have an analytic continuation off the diagonal in $`M\times \overline{M}`$ so that $`f(x,y)\psi (x,y)^l`$ is globally defined, smooth and bounded on $`K\times M`$ and on $`M\times K`$ for each compact subset $`K`$ of $`M`$ for some positive power $`l`$ and denote by $`_l`$ those for which the power $`l`$ suffices. Since $`\psi `$ is smooth and bounded it is clear that $``$ is a subalgebra of $`C^{\mathrm{}}(M)`$. In the case $`M`$ is compact we obviously have $`\widehat{E}(L^l)_l`$. If $`\widehat{A},\widehat{B}`$ belong to $``$, formula (13) is well defined for $`k`$ large enough.
We study the behaviour of the integral 13 in terms of $`k`$.
#### 4.1.3 An asymptotic formula
In order to localise the integral 13 we use a version of the Morse Lemma adapted from Combet, as in Moreno and Ortega-Navarro.
Let $`(M,\omega ,J)`$ be a Kรคhler manifold with metric $`g`$. We denote by $`\mathrm{exp}_xX`$ the exponential at $`x`$ of $`XT_xM`$. If $`g`$ is not complete the exponential map may not be defined for all $`x`$ and $`X`$, but in any case there is an open subset $`VTM`$ where it is defined and which contains the zero-section. The differential of the exponential map at $`0`$ is the identity so the map $`\alpha :VM\times M`$ given by $`\alpha (X)=(p(X),\mathrm{exp}_{p(X)}X)`$ where $`p`$ is the projection in the tangent bundle $`p:TMM`$ is a diffeomorphism near the zero-section. At any point of the zero-section the differential of $`\alpha `$ is the identity.
###### Proposition 33
Let $`(M,\omega ,J)`$ be a Kรคhler manifold with metric $`g`$ and $`\alpha :VM\times M`$ be the map defined above. Let $`(L,,h)`$ be a regular quantization bundle over $`M`$ and let $`\psi `$ be the corresponding $`2`$-point function on $`M\times M`$. Then there exists an open neighbourhood $`WV`$ of the zero-section in $`TM`$ and a smooth open embedding $`\nu :WTM`$ such that
$$(\mathrm{log}\psi \alpha \nu )(X)=\pi g_{p(X)}(X,X),XW$$
and the differential of $`\nu `$ at any point of the zero-section is the identity.
Denote by $`\stackrel{~}{}`$ the set of functions $`f`$ on $`M\times M\psi ^1(0)`$ such that $`f(x,y)\psi (x,y)^l`$ has a smooth extension to all of $`M\times M`$ which is bounded on $`K\times M`$ for each compact subset $`KM`$ for some $`l`$ and denote by $`\stackrel{~}{}_l`$ those for which the power $`l`$ suffices. If $`f`$, $`g`$ are in $``$ then $`f(x,y)g(y,x)`$ is in $`\stackrel{~}{}`$. Note also that if $`f\stackrel{~}{}`$ then its restriction to the diagonal $`\widehat{f}(x)=f(x,x)`$ is smooth.
For any $`f`$ belonging to $`\stackrel{~}{}_l`$, the integral
$$F_k(x)=_Mf(x,y)\psi (x,y)^kk^n\frac{\omega ^n(y)}{n!},\mathrm{for}\mathrm{k}\mathrm{l}+1$$
admits an asymptotic expansion
$$F_k(x)\underset{r0}{}k^rC_r(\widehat{f})(x)$$
where $`C_r`$ is a smooth differential operator of order $`2r`$ depending only on the geometry of $`M`$. The leading term is given by $`C_0(\widehat{f})(x)=\widehat{f}`$.
We are not claiming that for this very general class of functions $`f(x,y)`$ the integral depends smoothly on $`x`$, only that the coefficients of the asymptotic expansion do.
In the regular case $`ฯต^{(k)}`$ has an asymptotic expansion $`_{r0}ฯต_r/k^r`$ as $`k`$ tends to infinity with $`ฯต_0=1`$. Indeed,
$$1=1_k1=ฯต^{(k)}_M\psi (x,y)^kk^n\frac{\omega ^n(y)}{n!};$$
has an asymptotic expansion in $`k^1`$ with leading term $`1`$ by the previous proposition and we can then invert the asymptotic expansion to obtain one for $`ฯต^{(k)}`$.
###### Theorem 34
Let $`(M,\omega ,J)`$ be a Kรคhler manifold and $`(L,,h)`$ be a regular quantization bundle over $`M`$. Let $`\widehat{A},\widehat{B}`$ be in $``$. Then
$$(\widehat{A}_k\widehat{B})(x)=_M\widehat{A}(x,y)\widehat{B}(y,x)\psi ^k(x,y)ฯต^{(k)}k^n\frac{\omega ^n(y)}{n!}(y),$$
defined for $`k`$ sufficiently large, admits an asymptotic expansion in $`k^1`$ as $`k\mathrm{}`$
$$(\widehat{A}_k\widehat{B})(x)\underset{r0}{}k^rC_r(\widehat{A},\widehat{B})(x)$$
and the cochains $`C_r`$ are smooth bidifferential operators, invariant under the automorphisms of the quantization and determined by the geometry alone. Furthermore
$$C_0(\widehat{A},\widehat{B})=\widehat{A}\widehat{B},$$
and
$$C_1(\widehat{A},\widehat{B})C_1(\widehat{B},\widehat{A})=\frac{i}{\pi }\{\widehat{A},\widehat{B}\}.$$
#### 4.1.4 A convergent star product for flag manifolds
We first would like to show that the asymptotic expansion obtained above defines an associative formal star product. For this we assume that $`(M,\omega ,J)`$ is a flag manifold. Reshetikhin and Takhtajan have announced an analogous result for general Kรคhler manifolds.
Observe that if $`G`$ is a Lie group of isometries of the Kรคhler manifold $`(M,\omega ,J)`$ which lifts to a group of automorphisms of the quantization bundle $`(L,,h)`$ it acts naturally on $`(L^k,^{(k)},h^k)`$ :
$$g^{}(\widehat{A}_k\widehat{B})(x)=(g^{}\widehat{A}g^{}\widehat{B})(x)$$
for any $`\widehat{A},\widehat{B}`$ in $`\widehat{E}(L^l)`$ and any $`kl`$. Observe also that the bidifferential operators $`C_r`$ depend on the geometry alone thus are invariant under $`G`$.
###### Lemma 35
Let $`(M,\omega ,J)`$ be a flag manifold with $`M=G/K`$ where $`G`$ is a compact simply-connected Lie group and $`K`$ the centralizer of a torus. Assume the geometric quantization conditions are satisfied and let $`(L,,h)`$ be a quantization bundle over $`M`$. Let $`๐_L=_k\widehat{E}(L^k)`$ be the union of the symbol spaces. Then $`๐_L`$ coincides with the space $`E`$ of vectors in $`C^{\mathrm{}}(M)`$ whose $`G`$-orbit is contained in a finite dimensional subspace.
In the case of a flag manifold as above, the group $`G`$ lifts to a group of automorphisms of the quantization bundle $`(L^k,^{(k)},h^k)`$ hence the map $`\widehat{E}(L^l)\widehat{E}(L^l)C^{\mathrm{}}(M)`$ given by $`\widehat{A}\widehat{B}\widehat{A}_k\widehat{B}`$ intertwines the action of $`G`$ and the bidifferential operators $`C_r`$ are invariant under $`G`$. Thus, if $`\widehat{A}`$, $`\widehat{B}`$ belong to $`\widehat{E}(L^l)`$, there exists an integer $`a(l)`$ such that $`\widehat{A}_k\widehat{B}`$ belongs to $`\widehat{E}(L^{a(l)})`$ for all $`kl`$, and such that $`C_r(\widehat{A},\widehat{B})`$ belongs to $`\widehat{E}(L^{a(l)})`$ for every integer $`r`$.
Consider now the asymptotic development:
$$\widehat{A}_k\widehat{B}=\underset{r=0}{\overset{N}{}}k^rC_r(\widehat{A},\widehat{B})+R_N(\widehat{A},\widehat{B},k)$$
where
$$\underset{k\mathrm{}}{lim}k^NR_N(\widehat{A},\widehat{B},k)=0.$$
The above tells us that $`R_N(\widehat{A},\widehat{B},k)`$ belongs to $`\widehat{E}(L^{a(l)})`$ where $`a(l)`$ is independent of $`k`$. So, we can write
$`(\widehat{A}_k\widehat{B})_k\widehat{C}={\displaystyle \underset{r=0}{\overset{N}{}}}k^rC_r(\widehat{A},\widehat{B})_k\widehat{C}+R_N(\widehat{A},\widehat{B})_k\widehat{C}`$ (14)
$`={\displaystyle \underset{r,s=0}{\overset{N}{}}}k^{rs}C_s(C_r(\widehat{A},\widehat{B}),\widehat{C})`$ (15)
$`+{\displaystyle \underset{r=0}{\overset{N}{}}}k^rR_N(C_r(\widehat{A},\widehat{B}),\widehat{C},k)+R_N(\widehat{A},\widehat{B},k)_k\widehat{C}.`$ (16)
The last two terms multiplied by $`k^N`$ tend to zero when $`k`$ tends to infinity.
###### Theorem 36
The asymptotic expansion $`_{r0}k^rC_r(u,v)`$ yields a formal associative deformation of the usual product of functions in $`๐_L`$. It is a formal star product which extends to all of $`C^{\mathrm{}}(M)`$, using uniform convergence.
We prove that $`\widehat{A}_k\widehat{B}`$ is a rational function of $`k`$ with no pole at infinity, when the flag manifold is a hermitian symmetric space, by using structure theory of these spaces.
###### Theorem 37
Let $`M`$ be a compact hermitian symmetric space and let $`(L,,h)`$ be a quantization bundle over M. Let $`L^k=^kL`$ and let $`^k`$ be the space of holomorphic sections of $`L^k`$. Let $`\widehat{E}(L^k)`$ be the space of symbols of operators on $`^k`$. If $`\widehat{A}`$, $`\widehat{B}`$ belong to $`\widehat{E}(L^l)`$ and $`kl`$, the product $`\widehat{A}_k\widehat{B}`$ depends rationally on k and has no pole at infinity, hence the asymptotic expansion of $`\widehat{A}_k\widehat{B}`$ is convergent.
This result is generalised by Karabegov :
###### Theorem 38
For any generalised flag manifold the $`_k`$ product of two symbols is a rational function of $`k`$ without pole at infinity.
#### 4.1.5 Star product on bounded symmetric domains
Bounded symmetric domains
Let $`๐`$ denote a bounded symmetric domain. We shall use the Harish-Chandra embedding to realise $`๐`$ as a bounded subset of its Lie algebra of automorphisms. More precisely, if $`G_0`$ is the connected component of the group of holomorphic isometries then $`๐`$ is the homogeneous space $`G_0/K_0`$ where $`G_0`$ is a non-compact semi-simple Lie group and $`K_0`$ is a maximal compact subgroup. Let $`๐ค`$ be the Lie algebra of $`G_0`$, $`๐จ`$ the subalgebra corresponding with $`K_0`$, $`๐ค`$ and $`๐จ`$ the complexifications and $`G`$, $`K`$ the corresponding complex Lie groups containing $`G_0`$ and $`K_0`$. The complex structure on $`๐`$ is determined by $`K_0`$-invariant abelian subalgebras $`๐ช_+`$ and $`๐ช_{}`$ with
$$๐ค=๐ช_++๐จ+๐ช_{},[๐ช_+,๐ช_{}]๐จ,\overline{๐ช_+}=๐ช_{}$$
where $`\overline{}`$ denotes conjugation over the real form $`๐ค`$ of $`๐ค`$. The exponential map sends $`๐ช_\pm `$ diffeomorphically onto subgroups $`M_\pm `$ of $`G`$ such that $`M_+KM_{}`$ is an open set in $`G`$ containing $`G_0`$ and the multiplication map
$$M_+\times K\times M_{}M_+KM_{}$$
is a diffeomorphism. $`KM_{}`$ is a parabolic subgroup of $`G`$ and the quotient $`G/KM_{}`$ a generalised flag manifold. The $`G_0`$-orbit of the identity coset can be identified with $`G_0/K_0`$ and lies inside $`M_+KM_{}/KM_{}M_+`$. Composing this identification with the inverse of the exponential map gives the desired Harish-Chandra embedding of $`๐`$ as a bounded open subset of $`๐ช_+`$. We shall assume from now on that $`๐๐ช_+`$ via this embedding. In this realisation it is clear that the action of $`K_0`$ on $`๐`$ coincides with the adjoint action of $`K_0`$ on $`๐ช_+`$.
Following Satake, we define maps
$$k:๐\times ๐K,m_\pm :๐\times ๐๐ช_\pm $$
by
$$\mathrm{exp}\overline{Z^{}}\mathrm{exp}Z=\mathrm{exp}m_+(Z,Z^{})k(Z,Z^{})^1\mathrm{exp}m_{}(Z,Z^{}),$$
for $`Z,Z^{}๐.`$ They satisfy
$$\overline{k(Z,Z^{})}=k(Z^{},Z)^1,\overline{m_+(Z,Z^{})}=m_{}(Z^{},Z).$$
The holomorphic quantization
For any unitary character $`\chi `$ of $`K_0`$ there is a Hermitian holomorphic line bundle $`L`$ over $`๐`$ whose curvature is the Kรคhler form of an invariant Hermitian metric on $`๐`$. If $`\chi `$ also denotes the holomorphic extension to $`K`$ then the Hermitian metric has Kรคhler potential $`\mathrm{log}\chi (k(Z,Z))`$ and $`L`$ has a zero-free holomorphic section $`s_0`$ with
$$|s_0(Z)|^2=\chi (k(Z,Z))^1.$$
For $`\chi `$ sufficiently positive $`s_0`$ is square-integrable and the representation $`U`$ of $`G_0`$ on the space $``$ of square-integrable sections of $`L`$ is one of Harish-Chandraโs holomorphic discrete series. $`s_0`$ is a highest weight vector for the extremal $`K`$-type so is a smooth vector for the representation. We form the coherent states $`e_q`$ and see that $`e_{s_0(0)}`$ transforms the same way as $`s_0`$ and so they must be equal up to a multiple. This means that the coherent states are also smooth vectors of the representation. Further, since the quantization is homogeneous, $`ฯต`$ will be constant. Thus
$$e_{s_0(Z^{})},e_{s_0(Z)}=ฯต\chi (k(Z,Z^{})).$$
###### Lemma 39
Up to a constant (determined by the normalization of Haar measure on $`G_0`$) $`ฯต`$ is the formal degree $`d_U`$ of the discrete series representation $`U`$, hence by Harish-Chandraโs formula,it is a polynomial function of the differential $`d\chi `$ of the character $`\chi `$.
The two-point function $`\psi `$ is given by
$$\psi (Z,Z^{})=\frac{|\chi (k(Z,Z^{}))|^2}{\chi (k(Z,Z))\chi (k(Z^{},Z^{}))};$$
it takes the value 1 only on the diagonal.
Polynomial differential operators and symbols
We let $`๐`$ denote the algebra of holomorphic differential operators on functions on $`๐`$ with polynomial coefficients. We filter $`๐`$ by both the orders of the differentiation and the degrees of the coefficients: $`๐_{p,q}`$ denotes the subspace of operators of order at most $`p`$ with coefficients of degree at most $`q`$. Obviously, the composition of operators gives a map
$$๐_{p,q}\times ๐_{p^{},q^{}}๐_{p+p^{},q+q^{}}.$$
The global trivialization by $`s_0`$ of the holomorphic line bundle $`L`$ corresponding with the character $`\chi `$ allows us to transport the above operators to act on sections of $`L`$ by sending $`D๐`$ to $`D^\chi `$ where
$$D^\chi (fs_0)=(Df)s_0.$$
Let $`๐(\chi )`$ denote the resulting algebra of operators on sections of $`L`$ and $`๐_{p,q}(\chi )`$ the corresponding subspaces.
In this non-compact situation elements of $`๐(\chi )`$ do not define bounded operators on the Hilbert space $``$, but the fact that the coherent states are smooth vectors of the holomorphic discrete series representation and that polynomials are bounded on $`๐`$ means that each operator in $`๐(\chi )`$ maps the coherent states into $``$ so that it makes sense to speak of the symbols of these operators.
###### Lemma 40
The analytically continued symbol $`\widehat{D^\chi }(Z,Z^{})`$ of an operator $`D^\chi `$ in $`๐_{p,q}(\chi )`$ is a polynomial in $`Z`$ and $`m_{}(Z,Z^{})`$ of bidegree $`p,q`$.
The space of symbols of the operators in $`๐_{k,l}(\chi )`$ is the space of polynomials in $`Z`$ and $`m_{}(Z,Z^{})`$ of bidegree $`k,l`$ so, in particular, is independent of $`\chi `$.
Denote by $`_{p,q}`$ the space of polynomials in $`Z`$ and $`m_{}(Z,Z^{})`$ of bidegree $`p,q`$ and by $``$ the union of these spaces. $``$ is an algebra under pointwize multiplication, the algebra of symbols . If we take a symbol in $`_{p,q}`$ then it is the symbol of an operator in $`๐_{p,q}(\chi )`$. Taking two such operators and composing them corresponds with the composition of two polynomial operators in $`๐_{p,q}`$ and so can be expressed in terms of a basis for $`๐_{2p,2q}`$ as a rational function of $`d\chi `$. In other words the Berezin product $`fg`$ of two symbols $`f`$, $`g`$ in $`_{p,q}`$ is a symbol in $`_{2p,2q}`$ depending rationally on $`d\chi `$.
The star product
We construct a formal deformation of the algebra $`C^{\mathrm{}}(๐)`$ by first constructing it on the subalgebra $``$.
We consider the powers $`L^k`$ of the line bundle $`L`$ which correspond with the powers $`\chi ^k`$ of $`\chi `$. These powers have differentials $`kd\chi `$, so the Berezin product $`f_kg`$ of two symbols $`f`$, $`g`$ in $`_{p,q}`$ is a rational function of $`k`$ by the results of the previous section.
The symbols in $`_{p,q}`$ are in $`_0`$ (it is enough to show that $`adm_{}(Z,Z^{})`$ is bounded on $`X\times ๐`$ and $`๐\times X`$ for any compact subset $`X`$ of $`๐`$ ); thus the asymptotic expansion exists.
Since the Berezin product is associative for each $`k`$, the same argument as in the compact situation shows that its asymptotic expansion in $`k^1`$ is an associative formal deformation on $``$ with bidifferential operators as coefficients. To see that it extends to all of $`C^{\mathrm{}}(๐)`$ we show that $``$ contains enough functions to determine these operators, hence the asymptotic expansion of $`f_kg`$ has bidifferential operator coefficients which satisfy the cocycle conditions to define a formal product on $`C^{\mathrm{}}(๐)`$ which is associative.
###### Theorem 41
Let $`๐`$ be a bounded symmetric domain and $``$ the algebra of symbols of polynomial differential operators on a homogeneous holomorphic line bundle $`L`$ over $`๐`$ which gives a realisation of a holomorphic discrete series representation of $`G_0`$ (i.e $``$ is the algebra of functions on $`๐`$ which are polynomials in $`Z`$ and $`m_{}(Z,Z)),`$then for $`f`$ and $`g`$ in $``$ the Berezin product $`f_kg`$ has an asymptotic expansion in powers of $`k^1`$ which converges to a rational function of $`k`$. The coefficients of the asymptotic expansion are bidifferential operators which define an invariant and covariant star product on $`C^{\mathrm{}}(๐)`$. |
warning/0003/gr-qc0003086.html | ar5iv | text | # Inertial Effects on Berryโs Phase of Neutrino Oscillations
## 1 Introduction
Particle interferometry has been recognized as a tool to carry out delicate measurement of physical quantities that can be related to the difference in phase of the interferring beams. Besides, the observation of quantum mechanical phases provides accurate informations about quantities determining the phase shift (as for example, the flux of a magnetic field) and allows to test the quantum behaviour of systems. A remarkable contribution in this topic has been given by Berry in his pioneering paper. In his work , he showed that a quantum state acquires an additional phase (Berryโs phase) $`\beta ^g`$, which is related to the geometry of parameter space, besides the normal dynamical phase $`\varphi `$. He considered a nonโdegenerate quantum system in an initial eigenstate of a Hamiltonian which varies adiabatically through a circuit $`C`$ in the parametric space. The state evolves under the Schrรถdinger equation remaining in each instant an eigenstate of the Hamiltonian.
Aharonov and Anandan reformulated and generalized Berryโs result by disregarding the parameter space and considering an arbitrary, not necessarily adiabatic, cyclic evolution in the projective Hilbert space $`๐ซ()`$, which is the space of oneโdimensional subspace (called also rays) of an appropriate Hilbert space $``$ (see also ).
Nevertheless, in all formulations, Berryโs results are based on nonโrelativistic quantum mechanics and thus are not covariant. The covariant generalization of Berry phase was derived in Ref. where, by using the proper time method , the authors showed that
$`\beta ^g`$ $`=`$ $`i{\displaystyle _C}<\stackrel{~}{\mathrm{\Psi }}(\lambda )|{\displaystyle \frac{}{\lambda _a}}|\stackrel{~}{\mathrm{\Psi }}(\lambda )>๐\lambda _a`$
$`=`$ $`\varphi +i{\displaystyle _C}<\mathrm{\Psi }(\lambda )|{\displaystyle \frac{}{\lambda _a}}|\mathrm{\Psi }(\lambda )>๐\lambda _a,`$
where $`\lambda _a`$ are the evolution parameters depending on the proper time, the state $`\stackrel{~}{\mathrm{\Psi }}(\lambda )`$ satisfies the cyclic condition, $`|\stackrel{~}{\mathrm{\Psi }}(\mathrm{\Lambda }+\lambda )>=|\stackrel{~}{\mathrm{\Psi }}(\lambda )>`$, and $`|\mathrm{\Psi }(\mathrm{\Lambda }+\lambda )>=e^{i\varphi }|\mathrm{\Psi }(\lambda )>`$.
On the experimental side, a series of striking experiments have also been carried out and their results provide direct or indirect evidences for the existence of such a geometric phase .
The aim of this paper is to derive the geometrical phase attributed to the cyclic evolution of a mixed state, as for example, neutrino oscillations. This result is carried out in an nonโinertial reference frame, i.e. for an accelerating and rotating observer. As well known, the effects of the acceleration and the angular velocity are relevant in interferometry experiments. In fact, by using an accelerated neutron interferometer, Bonse and Wrobleski derived the predicted phase shift of quantum systems . Because of the validity of the equivalence principle, one expects that this effect occurs also in a gravitational field, as has been verified in Ref. . Besides, Mashhoon has derived a coupling of neutron spin to the rotation of a nonโinertial reference frame from an extension of the hypothesis of locality. In Ref. , the neutron Sagnac effect has been found using an angular velocity of about 30 times that of Earth. The spinโrotation and the spinโacceleration contributions to the helicity precession of fermions has been calculated in , and recently, the inertial effects on neutrino oscillations have been analyzed in .
We find that the geometrical phase, when evaluated in a nonโinertial reference frame, does depend on the parameters characterizing the physics of neutrino oscillations, i.e. the vacuum mixing angle and the massโsquared difference, and the ones characterizing the background geometry, the acceleration and angular velocity in our case. Besides, a dependence on the energy of neutrino also appear. The quantum phase shift induced by the non trivial topology of the background spaceโtime has, as we will see, interesting geometrical consequences in connection to the curvature of the parameter space of states.
The layout of this paper is the following. In Sect. 2, we review the main features of neutrino evolution in an accelerating and rotating frame (see Ref. ). The Berry phase in such a frame is calculated in Sect. 3. Discussion and Conclusions are drawn in Sect. 4.
## 2 Neutrino Evolution in a Non-Inertial Frame
In Ref. , the neutrino phase is generalized in the following way
$$|\psi (\lambda )>=\underset{j}{}U_{fj}e^{i_{\lambda _0}^\lambda Pp_{null}๐\lambda ^{}}|\nu _j>,$$
(2.1)
where flavor and mass indices are indicated by $`f`$ and by Latin letters, respectively. $`U_{fj}`$ are the matrix elements transforming flavor and mass bases, $`P`$ is the four momentum operator generating spaceโtime translation of the eigenstates and $`p_{null}^\mu =dx^\mu /d\lambda `$ is the tangent vector to the neutrino worldline $`x^\mu `$, parameterized by $`\lambda `$; $`x^\mu =(x^0,\stackrel{}{x})`$ are the local coordinates for the observer at the origin. We use the natural units.
The momentum operator $`P_\mu `$, used to calculate the phase of neutrino oscillations, is derived from the mass shell condition
$$(P_\mu +A_{G\mu }\gamma ^5)(P^\mu +A_G^\mu \gamma ^5)=M_f^2,$$
(2.2)
where $`A_G^\mu `$ are related to the spinorial connections appearing into the covariant Dirac equation in curved spaceโtime , $`M_f^2`$ is the vacuum mass matrix in flavor base
$$M_f^2=U\left(\begin{array}{cc}m_1^2& 0\\ 0& m_2^2\end{array}\right)U^{},U=\left(\begin{array}{cc}\mathrm{cos}\theta & \mathrm{sin}\theta \\ \mathrm{sin}\theta & \mathrm{cos}\theta \end{array}\right).$$
(2.3)
$`\theta `$ is the vacuum mixing angle. Ignoring terms of the order $`๐ช(A_G^2)`$ and $`๐ช(A_GM_f)`$, one gets that, for relativistic neutrinos moving along generic trajectories parameterized by $`\lambda `$, the column vector of flavor amplitude
$$\chi (\lambda )=\left(\begin{array}{c}<\nu _e|\psi (\lambda )>\\ <\nu _\mu |\psi (\lambda )>\end{array}\right)$$
(2.4)
satisfies the equation
$$i\frac{d\chi }{d\lambda }=\left(\frac{M_f^2}{2}+pA_G\gamma ^5\right)\chi .$$
(2.5)
In deriving (2.5), one uses the relation $`P^0=p^0`$ and $`P^ip^i`$ .
In a frame with acceleration $`\stackrel{}{a}`$ and angular velocity $`\stackrel{}{\omega }`$, the components of $`A_G^\mu `$ are
$$A_G^0=0,\stackrel{}{A}_G=\frac{\sqrt{g}}{2}\frac{1}{1+\frac{\stackrel{}{a}\stackrel{}{x}}{c^2}}\left\{2\frac{\stackrel{}{\omega }}{c}\frac{1}{c^2}[\stackrel{}{a}(\stackrel{}{x}\stackrel{}{\omega })]\right\},$$
(2.6)
so that Eq. (2.5) becomes
$$i\frac{d}{d\lambda }\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right)=๐ฏ\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right).$$
(2.7)
$`\sqrt{g}`$ is the determinant of the metric tensor of the accelerating and rotating frame , $`a_f<\nu _f|\psi (\lambda )>,f=e,\mu `$ and the matrix $`๐ฏ`$ is defined as
$$๐ฏ=\left[\begin{array}{cc}(\mathrm{\Delta }/2)\mathrm{cos}2\theta & (\mathrm{\Delta }/2)\mathrm{sin}2\theta \stackrel{}{p}\stackrel{}{A}_G\\ (\mathrm{\Delta }/2)\mathrm{sin}2\theta \stackrel{}{p}\stackrel{}{A}_G& (\mathrm{\Delta }/2)\mathrm{cos}2\theta \end{array}\right].$$
(2.8)
up to the $`(m_1^2+m_2^2)/2`$ term, proportional to the identity matrix. Here $`\mathrm{\Delta }(m_2^2m_1^2)/2`$. We restrict to flavors $`e,\mu `$, but this analysis works also for different neutrino flavors. In order to determine the mass eigenstates $`|\nu _1>`$ and $`|\nu _2>`$, one has to diagonalize the matrix $`๐ฏ`$. Using the standard procedure, one writes the mass eigenstates as a superposition of flavor eigenstates
$`|\nu _1>=\mathrm{cos}\stackrel{~}{\theta }|\nu _e>\mathrm{sin}\stackrel{~}{\theta }|\nu _\mu >,`$ (2.9)
$`|\nu _2>=\mathrm{sin}\stackrel{~}{\theta }|\nu _e>+\mathrm{cos}\stackrel{~}{\theta }|\nu _\mu >,`$
where the mixing angle $`\stackrel{~}{\theta }`$ is defined in terms of the vacuum mixing angle $`\theta `$
$$\mathrm{tan}2\stackrel{~}{\theta }=\frac{\mathrm{\Delta }\mathrm{sin}2\theta 2\stackrel{}{p}\stackrel{}{A}_G}{\mathrm{\Delta }\mathrm{cos}2\theta }.$$
(2.10)
We note that $`\stackrel{~}{\theta }\theta `$ as $`\stackrel{}{A}_G0`$ (i.e. $`\stackrel{}{a}0,\stackrel{}{\omega }0`$). The eigenvalues of the matrix (2.8) are
$$\tau _{1,2}=\pm \sqrt{\frac{\mathrm{\Delta }^2}{4}\mathrm{cos}^22\theta +\left[\frac{\mathrm{\Delta }}{2}\mathrm{sin}2\theta (\stackrel{}{p}\stackrel{}{A}_G)\right]^2}.$$
(2.11)
In the base of the mass eigenstates, we have $`|\psi (\lambda )>=a_1(\lambda )|\nu _1>+a_2(\lambda )|\nu _2>`$, so that Eq. (2.7) assumes the form
$$i\frac{d}{d\lambda }\left(\begin{array}{c}a_1\\ a_2\end{array}\right)=\left(\begin{array}{cc}\tau _1& 0\\ 0& \tau _2\end{array}\right)\left(\begin{array}{c}a_1\\ a_2\end{array}\right),$$
(2.12)
where $`a_i=<\nu _i|\psi (\lambda )>,i=1,2`$, and
$$\left(\begin{array}{c}a_1\\ a_2\end{array}\right)=\stackrel{~}{U}\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right),\stackrel{~}{U}=\left(\begin{array}{cc}\mathrm{cos}\stackrel{~}{\theta }& \mathrm{sin}\stackrel{~}{\theta }\\ \mathrm{sin}\stackrel{~}{\theta }& \mathrm{cos}\stackrel{~}{\theta }\end{array}\right).$$
(2.13)
We have used the adiabatic condition $`d\stackrel{~}{\theta }/d\lambda 0`$ in order that (2.12) is a diagonal matrix, i.e., we are neglecting the variations of acceleration and angular velocity, with respect to the affine parameter $`\lambda `$, in comparing to their magnitude. Eq. (2.12) implies
$$a_k(\lambda )=a_k(\lambda _0)e^{\alpha _k(\lambda )},\alpha _k(\lambda )i_{\lambda _0}^\lambda \tau _k๐\lambda ^{}.$$
(2.14)
with $`k=1,2`$. For the initial condition $`|\psi (\lambda _0)>=|\nu _e>`$, the state $`|\psi (\lambda )>`$ is
$`|\psi (\lambda )>`$ $`=`$ $`[\mathrm{cos}\theta _0\mathrm{cos}\stackrel{~}{\theta }e^{i\tau _1\lambda }+\mathrm{sin}\theta _0\mathrm{sin}\stackrel{~}{\theta }e^{i\tau _2\lambda }]|\nu _e>`$
$`+`$ $`[\mathrm{cos}\theta _0\mathrm{sin}\stackrel{~}{\theta }e^{i\tau _1\lambda }+\mathrm{sin}\theta _0\mathrm{cos}\stackrel{~}{\theta }e^{i\tau _2\lambda }]|\nu _\mu >,`$
where $`\theta _0=\stackrel{~}{\theta }(\lambda _0)`$ and the adiabaticity condition has been used in (2.14). Therefore, accelerating and rotating observers will experience a flavor oscillation of neutrinos. For example, the probability to observe an electron neutrino is readily calculated from (2) and it is given by
$$|<\nu _e|\psi (\lambda )>|^2=\mathrm{cos}^2(\theta _0+\stackrel{~}{\theta })\mathrm{sin}^2\alpha +\mathrm{cos}^2(\theta _0\stackrel{~}{\theta })\mathrm{cos}^2\alpha .$$
(2.16)
## 3 Inertial Effects on Berryโs Phase
In this Section, the Berry phase is calculated in an accelerating and rotating frame. For convenience, let us invert Eq. (2.9)
$`|\nu _e>`$ $`=`$ $`\mathrm{cos}\stackrel{~}{\theta }|\nu _1>+\mathrm{sin}\stackrel{~}{\theta }|\nu _2>,`$ (3.1)
$`|\nu _\mu >`$ $`=`$ $`\mathrm{sin}\stackrel{~}{\theta }|\nu _1>+\mathrm{cos}\stackrel{~}{\theta }|\nu _2>.`$
Repeating the previous analysis for the state $`|\nu _e(\lambda )>`$, we see that Eq. (2.7) can be read, formally, as a Schรถdinger equation, the matrix $`๐ฏ`$ playing the role of the effective Hamiltonian. Then, the neutrino state $`|\nu _e>`$ evolves according to the relation
$`|\nu _e(\lambda )>`$ $`=`$ $`e^{i๐ฏ\lambda }|\nu _e>`$ (3.2)
$`=`$ $`e^{i\tau _1\lambda }\mathrm{cos}\stackrel{~}{\theta }|\nu _1>+e^{i\tau _2\lambda }\mathrm{sin}\stackrel{~}{\theta }|\nu _2>,`$
being $`\tau _1`$ and $`\tau _2=\tau _1`$, defined in (2.11), the eigenvalues of $`๐ฏ`$ corresponding to the eigenkets $`|\nu _1>`$ and $`|\nu _2>`$, respectively (see Eq. (2.12)). In a cyclic evolution, i.e. from $`\lambda `$ to $`\mathrm{\Lambda }+\lambda `$, the final state $`|\nu _e(\mathrm{\Lambda }+\lambda )>`$ and the initial state $`|\nu _e(\lambda )>`$ (see Eq. (3.2)) do coincide, apart a phase factor, provided that $`\mathrm{\Lambda }=2\pi /(\tau _1\tau _2)`$
$$|\nu _e(\mathrm{\Lambda }+\lambda )>=e^{i\varphi }|\nu _e(\lambda )>,\varphi =\frac{2\pi \tau _1}{\tau _1\tau _2}=\pi ,$$
(3.3)
Inserting Eq. (3.3) into Eq. (1) (for $`a=1`$, i.e. the evolution parameter does coincide with the affine parameter), the Berry phase of an electron neutrino moving along a cyclic path can be written as
$$\beta _e^g=\varphi +i_0^\mathrm{\Lambda }<\nu _e(\lambda )|\frac{}{\lambda }|\nu _e(\lambda )>,$$
(3.4)
where $`\varphi `$ is the dynamical phase evaluated in Eq. (3.3). An explicit calculation of (3.4) yields the result
$`\beta _e^g`$ $`=`$ $`2\pi \mathrm{sin}^2\stackrel{~}{\theta }`$
$`=`$ $`\pi \left[1{\displaystyle \frac{\mathrm{\Delta }\mathrm{cos}2\theta }{\sqrt{\mathrm{\Delta }^2\mathrm{cos}^22\theta +(\mathrm{\Delta }\mathrm{sin}2\theta 2\stackrel{}{p}\stackrel{}{A}_G)^2}}}\right],`$
where Eqs. (2.10) and (3.1) have been used. In similar way we get the Berry phase of muon neutrino
$`\beta _\mu ^g`$ $`=`$ $`2\pi \mathrm{cos}^2\stackrel{~}{\theta }`$
$`=`$ $`\pi \left[1+{\displaystyle \frac{\mathrm{\Delta }\mathrm{cos}2\theta }{\sqrt{\mathrm{\Delta }^2\mathrm{cos}^22\theta +(\mathrm{\Delta }\mathrm{sin}2\theta 2\stackrel{}{p}\stackrel{}{A}_G)^2}}}\right],`$
As arises from Eqs. (3) and (3), $`\beta _{e,\mu }^g`$ depend on the vacuum mixing angle, on the massโsquared difference and energy of neutrinos, and on the parameters characterizing, through the spinorial connection, the background geometry: the acceleration and angular velocity. Besides, the quantum shift phase induces also a dependence on the energy of neutrinos. Note that $`\beta _e^g+\beta _\mu ^g=2\pi `$.
A direct or indirect evidence of the Berry phases (3) and (3) is very difficult to achieve and how a quantum interferometry experiment could reveal it goes beyond the aim this paper. We just analyze some interesting consequences of the above results.
If the condition
$$\mathrm{\Delta }\mathrm{sin}2\theta 2\stackrel{}{p}\stackrel{}{A}_G=0$$
(3.7)
holds, then Eqs. (2.10), (3) and (3) imply $`\stackrel{~}{\theta }=n\pi /2`$, $`n=0,1,2,\mathrm{}`$, so that
$$\beta _e^g=0\text{and}\beta _\mu ^g=2\pi .$$
(3.8)
Furthermore, in the rotating frame with angular velocity $`\omega `$ and acceleration zero, Eq. (3.7) reduces to the form
$$|m_2^2m_1^2|\frac{4E_\nu \omega }{\mathrm{sin}2\theta }$$
(3.9)
where $`E_\nu `$ is the energy of ultrarelativistic neutrinos, $`E_\nu |\stackrel{}{p}|`$. Let us analyze Eq. (3.9) for solar neutrinos and in the case in which the observer is comoving with the Earth, i.e. its angular velocity is $`\omega 710^5`$rad/sec. Results of the massโsquared difference, calculated by using (3.9) for typical values of the solar neutrino energies and vacuum mixing angle, are reported in the Table. The agreement with the experimental data comes from neutrinos with energy varying in the range $`10รท60`$MeV. In this range in fact, we find a massโsquared difference of the order $`10^{12}รท10^{10}`$eV<sup>2</sup> for vacuum mixing angle $`10^1\mathrm{sin}2\theta 1`$ (see also ).
At the present, there is a strong evidence in favor of oscillations of neutrinos and of their nonโzero masses. Such results have been found in different experiments: solar neutrino experiments , atmospheric neutrino experiments , and the accelerator LSND experiment . Recent reports indicate that the best fit in favor of solar neutrino oscillations are obtained for the following case
$$|m_2^2m_1^2|(0.5รท1.1)10^{10}\text{eV}\text{2},\mathrm{sin}^22\theta 0.67รท1,$$
in very good agreement with our results summarized in the Table.
If the momentum of the neutrino $`\stackrel{}{p}`$ is perpendicular to the spinorial field $`\stackrel{}{A}_G`$, then the coupling term vanishes and Eq. (3) and (3) reduce to
$$\beta _e^g=2\pi \mathrm{sin}^2\theta ,\beta _\mu ^g=2\pi \mathrm{cos}^2\theta ,$$
(3.10)
i.e. the mixing angle $`\stackrel{~}{\theta }`$ coincides with the vacuum mixing angle $`\theta `$. Of course, when the angular velocity and the acceleration of the frame are zero, we recover trivially Eq. (3.10) and results of Ref. , obtained for Minkowskian background. Eqs. (3) and (3) represent a generalization of the results obtained in Ref. .
For ultra-relativistic neutrinos and high values of the acceleration and angular velocity, so that the coupling term is much more greater than the massโsquared difference, the Berry phase (3) and (3) assume the following values: $`\beta _e^g\pi `$ and $`\beta _\mu ^g\pi `$.
Extension of these results can be done, invoking the equivalence principle, to stationary gravitational fields, last ones responsible of the quantum phase shift. The previous analysis applies also to quantum system involving boson mixing, as the oscillations of $`K^0\overline{K}^0`$.
## 4 Discussion and Conclusions
In this paper, the covariant generalization of the Berry phase proposed in Ref. has been applied to problems involving mixed states, as neutrino oscillations. We have calculated the Berry phase of the electric and muon neutrinos in a cyclic evolution with respect to an accelerating and rotating reference frame and we have found that these phases do depend on the vacuum mixing angle, the massโsquared difference and on the coupling between the momentum of neutrinos and the spinorial connection.
The condition of noโmixing expressed by Eq. (3.7), which diagonalizes the mixing matrix (2.8), implies that the Berry phase of the electron and muon neutrinos assume the values equal to $`0`$ or $`2\pi `$, respectively. Such a condition has been analyzed for observers comoving with the Earth, so that $`\omega 10^5`$rad/sec and $`|\stackrel{}{a}|=0`$. In this particular case, the numerical values of the massโsquared difference, calculated by using Eq. (3.9) (see Table), are compared with ones of the recent experimental data, showing a very good agreement. It is worth to note that the inertial effects on the Berry phase, hence on the neutrino oscillations, induce a dependence on the orientation of the rotating observer (the direction of the angular velocity) with respect to the direction of the neutrino momenta. It implies hence a dependence on the zenith angle recently discussed in the framework of the atmospheric neutrino flux .
A direct or indirect measurement of inertial (or gravitational) effects on the Berry phases related to the neutrino particles is at the moment very difficult, considering also the fact that the values of the massโsquared difference of neutrinos and the vacuum mixing angle are till now open issues. Nevertheless, the Berry phases (3) and (3) have interesting implications from the geometrical point of view. As already said in the Introduction, the Berry phase is attributed to the holonomy in the parameter space. It appears when a quantum state is parallelly transported in the parameters space around a close path. The final quantum state is rotated, with respect to the initial one, by an angle which can be related to the connection (the Berry connection) $`๐=<\stackrel{~}{\mathrm{\Psi }}|d|\stackrel{~}{\mathrm{\Psi }}>=๐_ad\lambda ^a`$ appearing in Eq. (1), being $`d=(/\lambda _a)d\lambda ^a`$ the exterior derivative in the parameter space. Besides, one can also define a curvature (the Berry curvature) as the field strength of $`๐`$, $`=d๐=(d<\stackrel{~}{\mathrm{\Psi }}|)(d|\stackrel{~}{\mathrm{\Psi }}>)`$.
The existence of this phase, associated with the cyclic evolution, is universal in the sense that it is the same for the infinite number of possible motions along the curves in the Hilbert space, which project to a given closed curve in the projective Hilbert space of the rays, and for the possible (effective) Hamiltonians, which rule the evolution of the state along these curves.
In a Minkowski spaceโtime, the Berry phase is related only to the vacuum mixing angle: The quantum vector state is rotated of the angle $`\theta `$ when parallelly transported along a close curve, so that, as discussed above, it can be associated to the curvature in the space of states (see Eq. (3.10) and Ref. ).
In a nonโinertial reference frame, new geometrical features arise, as one can see from Eqs. (3) and (3):
The parallel transport along a close curve contains also the contribution due to the spinorial connection $`A_G^\mu `$, which induces a quantum shift phase, hence an additional rotation in the parameter space of the quantum vector state. The total rotation is given by the (mixing) angle $`\stackrel{~}{\theta }`$, so that the Berry curvature turns out to be related both the vacuum mixing angle and to the spinorial connection, i.e. to the non trivial geometry of the background spaceโtime.
Due to the coupling term, $`\stackrel{}{p}\stackrel{}{A}_G`$, the geometrical phases (3) and (3) show a dependence on the energy of the neutrinos, loosing in such a way the character of universality: different values of the neutrino energies correspond to different Berryโs phase.
Such a new geometrical setting follows from the generalization of the quantum mechanical phase (2.1) in which, as suggested by Stodolsky , the metric field appears in the definition of the scalar product. The dependence of the geometrical phase on the energy of particles and on the parameters characterizing the background geometry is a common aspect of the behaviour of a quantum particle in a certain class of four dimensional stationary spaceโtime, as tubular matter source, slowly moving mass current and spinning cosmic string . In the last case, for example, it is shown that the gravitational geometrical phase of a quantum particle with energy $`E`$ moving around a spinning cosmic string (with angular velocity $`J`$) is $`JE`$ .
As final comment, we stress that the previous results holds also for stationary gravitational field, as an obvious consequence of the validity of the equivalence principle. Then, the understanding of the issues concerning the influence of the gravity on the Berry phase are of interest in view of theoretical arguments which try to construct general relativity as a gauge theory and, in particular, to formulate the symmetric secondโrank tensor field in the framework of the unified gauge approach. In such a way, the gravitational field should play the role of electromagnetic field, in strict analogy with the BohmโAharonov effect, and its theoretical (and phenomenological) consequence could help for a better understanding of the quantum nature of particles evolving in non trivial geometries.
Table: Estimation of $`|m_2^2m_1^2|`$ given by (3.9) as function of $`E_\nu `$, $`\mathrm{sin}2\theta `$, and fixed value of $`\omega 710^5`$rad/sec.
| $`E_\nu `$(MeV) | $`\mathrm{sin}2\theta `$ | $`|m_2^2m_1^2|`$(eV<sup>2</sup>) |
| --- | --- | --- |
| 1 | 1 | $`10^{13}`$ |
| 1 | $`10^1`$ | $`10^{12}`$ |
| 10 | 1 | $`10^{12}`$ |
| 10 | $`10^1`$ | $`10^{11}`$ |
| $`50รท60`$ | 1 | $`10^{10}`$ |
Acknowledgments
Research supported by MURST fund 40% and 60% art. 65 D.P.R. 382/80.
GL acknowledges UE (P.O.M. 1994/1999) for financial support. |
warning/0003/hep-ph0003150.html | ar5iv | text | # Random matrix model for chiral symmetry breaking and color superconductivity in QCD at finite density.An animated gif movie showing the evolution of the phase diagram with the chiral and diquark coupling parameters can be viewed at http://www.nbi.dk/~vdheyden/QCDpd.html
## I Introduction
A number of early and recent model studies of finite density QCD have suggested that quark Cooper pairs may form above some critical density and lead to โcolor superconductingโ matter. Although perturbation theory performed on single-gluon exchange suggests pairing gaps of a few MeV , some recent calculations including non-perturbative interactions, either in the form of the Nambu Jona-Lasinio (NJL) model or those induced by instantons , indicate gaps as large as $`100`$ MeV. Theses values imply that color superconductivity may be relevant to the physics of heavy-ion collisions and neutron stars.
Quark pairing singles out one color direction and thus spontaneously breaks $`SU(3)_{\mathrm{color}}`$ to $`SU(2)_{\mathrm{color}}`$. The pattern of symmetry breaking may however be richer, since the formation of condensates in one $`qq`$ channel competes with the breaking of chiral symmetry in the orthogonal $`\overline{q}q`$ channel. In an earlier paper we formulated random matrix models for both chiral and diquark condensations in the limit of two quarks flavors and zero chemical potential. Our aim was to understand the phase structures which result from the competition between the two forms of order solely on the basis of the underlying symmetries. In this spirit, we constructed random matrix interactions for which the single quark Hamiltonian satisfies two basic requirements: (1) its block structure reflects the color $`SU(N_c)`$ and chiral $`SU(2)_\mathrm{L}\times SU(2)_\mathrm{R}`$ symmetries of QCD and (2) the single quark Hamiltonian is Hermitian. This last condition ensures the existence of well-defined relationships between the order parameters and the spectral properties of the interactions. In particular, condition (2) is obeyed by single-gluon exchange, which is generally regarded as the relevant description of QCD. Aside from conditions (1) and (2), the dynamics of the interactions does not contain any particular structure. In order to solve the model exactly, we described this โdynamicsโ by independent Gaussian distributions of matrix elements.
In practice, what distinguishes the interactions from one another is their respective coupling constants, $`A`$ and $`B`$, in the $`qq`$ and $`\overline{q}q`$ channels. The ratio $`B/A`$ measures the balance between chiral and diquark condensation forces. The absolute magnitudes of $`A`$ and $`B`$ play a secondary role. They introduce a scale in the condensation fields but do not affect the phase structure. We have shown in that the condition (2) of Hermiticity forces $`B/A`$ to be smaller or equal to $`N_c/2`$. This constraint results in the absence of a stable diquark phase in the limit of zero chemical potential.
The purpose of this paper is to extend our previous analysis to non-zero chemical potentials, for which the interactions cease to be Hermitian. Non-Hermitian interactions lead to considerable difficulties in numerical calculations in both lattice and random matrix theories. Standard Monte Carlo techniques fail because the fermion determinant in the action is complex and the sampling weights are no longer positive definite. Obtaining reliable results thus requires a proper treatment of cancellations between a large number of terms; a problem which has not yet found a satisfactory solution . The present random matrix models possess exact solutions. The saddle-point methods used in to derive the free-energy as a function of the condensation fields remain valid at finite densities and give exact results in the thermodynamic limit of matrices of infinite dimensions. Here, we will use these methods to calculate the thermodynamics quantities as a function of the condensation fields and deduce the phase diagrams from the field configurations which maximize the pressure.
We will discover that the pressure function has a simple analytic dependence which leads to polynomial gap equations. This situation reminds us of the gap equations obtained in a Landau-Ginzburg theory near criticality, and the random matrix approach is analogous in many respects. Both theories associate a change of symmetry with a change of state in the system. They are both mean-field and describe the dynamics of a reduced number of degrees of freedom $`N`$, where $`N`$ scales with the volume of the system. In random matrix models, these degrees of freedom can be related to the low-lying quark excitation modes in the gluon background, i.e. the zero modes in an instanton approximation of that background <sup>*</sup><sup>*</sup>*This relation is explicit in the random matrix models where only chiral symmetry is considered ; it is expected to remain true once color symmetry is also taken into account .. In the Landau-Ginzburg formulation, the degrees of freedom correspond to the long-wavelength modes which remain after coarse-graining. There is, however, an essential difference in the construction of the two theories. A Landau-Ginzburg theory starts with the specification of an effective potential for the relevant degrees of freedom. A random matrix model starts at a more microscopic level with the construction of an interaction which, once integrated over, produces an effective potential. The integration can carry along dynamical constraints and thus restricts the allowed range of coupling constants. An example of such restriction is the Hermiticity condition in , which implies that $`B/AN_c/2`$. This condition is characteristic of the dynamics of the interactions which we consider and remains true at finite chemical potential. We will therefore take it into account in the following.
In this paper, we consider random matrix models in which both chiral and color condensations can take place and study the resulting phase diagrams in temperature and chemical potential. We review the model of Ref. and discuss the general form of the pressure as a function of the condensation fields in Sec. II. We then solve the gap equations and analyze the six topologies that the phase diagram can assume in Sec. III. We compare our results with those of QCD effective models, discuss the possibility of other symmetry breaking patterns and the extension to the case of non-zero quark current masses in Sec. IV. Section V presents our conclusions.
## II Formulation of the model
### A The partition function
The generalization of the matrix models introduced in Ref. to finite quark densities is straightforward. We represent the quark fields for each of the two flavors by $`\psi _1^{}`$ and $`\psi _2^{}`$. The partition function at temperature $`T`$ and chemical potential $`\mu `$ is then
$`Z(\mu ,T)={\displaystyle ๐H๐\psi _1^{}๐\psi _1^{}๐\psi _2^T๐\psi _2^{}}`$ (1)
$`\times \mathrm{exp}\left[i\left(\begin{array}{c}\psi _1^{}\\ \text{}\\ \psi _2^T\end{array}\right)^T\left(\begin{array}{cc}+\left(\pi T+i\mu \right)\gamma _0+im& \eta P_\mathrm{\Delta }\\ \text{}& \text{}\\ \eta ^{}P_\mathrm{\Delta }& ^T+\left(\pi Ti\mu \right)\gamma _0^Tim\end{array}\right)\left(\begin{array}{c}\psi _1^{}\\ \text{}\\ \psi _2^{}\end{array}\right)\right].`$ (11)
Here, $``$ is a matrix of dimension $`4\times N_c\times N`$ which represents the interaction of a single quark with a gluon background. Its measure, $`๐H`$, will be discussed below. The parameters $`m`$ and $`\eta `$ select a particular direction for chiral and color symmetry breaking and are to be taken to zero in the appropriate order at the end of the calculations. The current quark mass $`m`$ is to be associated with the chiral order parameters $`\psi _1^{}\psi _1^{}`$ and $`\psi _2^T\psi _2^{}`$. The complex parameter $`\eta `$ is to be associated with the order parameter for Cooper pairing, $`\psi _2^TP_\mathrm{\Delta }\psi _1^{}`$, in which $`P_\mathrm{\Delta }iC\gamma _5\lambda _2`$ ($`C`$ is the charge conjugation operator) selects the quark-quark combinations which are antisymmetric in spin and color, i.e. a chiral isosinglet, Lorentz scalar, and color $`\overline{3}`$ state. Note that in order to permit the construction of correlations between the fields $`\psi _1`$ and $`\psi _2^T`$, we have transposed the single quark propagator in the second flavor, hence the upperscript $`T`$.
The interaction $``$ is intended to mimic the effects of gluon fields and thus explicitly includes the desired chiral and color symmetries. Of central interest is single-gluon exchange, which has the chiral block-structure
$`_{\mathrm{sge}}`$ $`=`$ $`\left(\begin{array}{cc}0& W_{\mathrm{sge}}\\ W_{\mathrm{sge}}^{}& 0\end{array}\right),`$ (14)
where $`W_{\mathrm{sge}}`$ has the spin and color block-structure of a vector interaction,
$`W_{\mathrm{sge}}`$ $`=`$ $`{\displaystyle \underset{\mu =1}{\overset{4}{}}}{\displaystyle \underset{a=1}{\overset{8}{}}}\sigma _\mu ^+\lambda ^aA^{\mu a}.`$ (15)
Here, $`\sigma _\mu ^+=(1,i\stackrel{}{\sigma })_\mu `$ are the $`2\times 2`$ spin matrices, and $`\lambda _a`$ denote the $`N_c\times N_c`$ Gell-Mann matrices. The $`A^{\mu a}`$ are real $`N\times N`$ matrices which represent the gluon fields.
Since we want to explore the evolution of the phase structure as the balance between chiral and diquark condensations is changed, we consider the larger class of Hermitian interactions to which single-gluon exchange belongs. As noted in the introduction, this choice is motivated by the fact that Hermitian interactions have a clear relationship between the order parameters and the spectral properties, in the form of Banks-Casher formulae . We write an Hermitian interaction $``$ as an expansion into a direct product of the sixteen Dirac matrices $`\mathrm{\Gamma }_C`$ times the $`N_c^2`$ color matrices. The matrix elements are given by
$`_{\lambda i\alpha k;\kappa j\beta l}`$ $`=`$ $`{\displaystyle \underset{C=1}{\overset{16}{}}}\left(\mathrm{\Gamma }_C^{}\right)_{\lambda i;\kappa j}{\displaystyle \underset{a=1}{\overset{N_c^2}{}}}\mathrm{\Lambda }_{\alpha \beta }^a\left(A_{\lambda \kappa }^{Ca}\right)_{kl},`$ (16)
where the indices ($`\lambda `$,$`\kappa `$), $`(i,j)`$, and $`(\alpha ,\beta )`$ respectively denote chiral, spin, and color quantum numbers, while $`(k,l)`$ are matrix indices running from $`1`$ to $`N`$. The $`\mathrm{\Lambda }^a`$ represent the color matrices $`\lambda ^a`$ when $`aN_c^21`$ and the diagonal matrix $`(\delta _c)_{\alpha \beta }=\delta _{\alpha \beta }`$ when $`a=N_c^2`$. The normalization for color matrices is $`\mathrm{Tr}[\lambda ^a\lambda ^b]=2\delta _{ab}`$ and $`\mathrm{Tr}[\delta _c^2]=N_c`$; the normalization of the Dirac matrices is $`\mathrm{Tr}[\mathrm{\Gamma }_C\mathrm{\Gamma }_C^{}]=4\delta _{CC^{}}`$.
The random matrices $`A^{Ca}`$ are real when $`C`$ is vector or axial vector ($`C=V,A`$) and real symmetric when $`C`$ is scalar, pseudoscalar, or tensor $`(C=S,P,T)`$ . Their measure is
$`๐H`$ $`=`$ $`\left\{{\displaystyle \underset{Ca}{}}{\displaystyle \underset{\lambda \kappa }{}}๐A_{\lambda \kappa }^{Ca}\right\}\mathrm{exp}\left[N{\displaystyle \underset{Ca}{}}{\displaystyle \underset{\lambda \kappa }{}}\beta _C\mathrm{\Sigma }_{Ca}^2\mathrm{Tr}[A_{\lambda \kappa }^{Ca}(A_{\lambda \kappa }^{Ca})^T]\right],`$ (17)
where $`๐A_{\lambda \kappa }^{Ca}`$ are Haar measures. Here, $`\beta _C=1`$ for $`C=V,A`$ and $`\beta _C=1/2`$ for $`C=S,P,T`$. We want to mimic interactions which in a four-dimensional field theory would respect color $`SU(N_c)`$ and Lorentz invariance in the vacuum. Therefore, we choose a single variance $`\mathrm{\Sigma }_{Ca}`$ for all channels which transform equally under color and space rotations.
The temperature and chemical potential enter the model in Eq. (23) through the inclusion of the first Matsubara frequency in the single quark propagator. Such $`T`$ and $`\mu `$ dependence is certainly oversimplified but none the less sufficient to produce the desired physics. Our purpose is to understand the general topology of the phase diagram and not to provide explicit numbers. More refined treatments including, for instance, all Matsubara frequencies would modify the details of the phase diagram and map every $`(\mu ,T)`$ coordinate to a new one. However, any such mapping will necessarily be monotonic and will conserve the topology. We note in particular that we do not assume a temperature dependence of the variances, i.e. we neglect the $`T`$-dependence of the gluon background. A non-analytic behavior in $`T`$ should only arise in the contribution to the thermodynamics from the degrees of freedom related to chiral and diquark condensations. We thus expect a realistic $`T`$-dependence of the gluon background (and of the variances) to be smooth and not to affect the overall phase topology.
We choose the signs of the $`T`$\- and $`\mu `$-dependencies to mimic a diquark condensate which is uniform in time, i.e. which does not contain a proper pairing frequency. In a microscopic theory formulated in four-momentum space, the absence of a proper frequency leads to pairing between particle and hole excitations with energies which are symmetric around the Fermi surface. We simulate this effect by selecting opposite $`T`$-dependences for the fields $`\psi _1^{}`$ and $`\psi _2^T`$, while maintaining the same $`\mu `$-dependence.
### B The pressure function
The integration over the random matrix interactions is Gaussian and can thus be performed exactly. Following the procedure of Ref. , we use a Hubbard-Stratonovitch transformation to introduce two auxiliary variables $`\sigma `$ and $`\mathrm{\Delta }`$, to be associated with the chiral and pairing order parameters respectively
$`\sigma `$ $``$ $`\psi ^{}\psi \psi _1^{}\psi _1^{}=\psi _2^T\psi _2^{},`$ (18)
$`\mathrm{\Delta }`$ $``$ $`\psi \psi \psi _2^TP_\mathrm{\Delta }\psi _1^{}=\left(\psi _1^{}P_\mathrm{\Delta }\psi _2^{}\right)^{}.`$ (19)
An integration over the fermion fields then reduces the partition function to
$`Z(\mu ,T)`$ $`=`$ $`{\displaystyle ๐\sigma ๐\mathrm{\Delta }\mathrm{exp}\left[4N\mathrm{\Omega }(\sigma ,\mathrm{\Delta })\right]},`$ (20)
where $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ is the negative of the pressure, $`P(\sigma ,\mathrm{\Delta })`$, per degree of freedom and per unit spin and flavor,
$`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ $`=`$ $`P(\sigma ,\mathrm{\Delta })`$ (21)
$`=`$ $`A\mathrm{\Delta }^2+B\sigma ^2{\displaystyle \frac{1}{2}}\{(N_c2)\mathrm{log}\left[((\sigma +m\mu )^2+T^2)((\sigma +m+\mu )^2+T^2)\right]`$ (23)
$`+2\mathrm{log}\left[((\sigma +m\mu )^2+T^2+|\mathrm{\Delta }+\eta |^2)((\sigma +m+\mu )^2+T^2+|\mathrm{\Delta }+\eta |^2)\right]\},`$
where we have dropped the prefactors $`\pi `$ in the temperature dependence for simplicity. Here, the coupling constants $`A`$ and $`B`$ are weighted averages of the Fierz coefficients $`f_\mathrm{\Delta }^{Ca}`$ and $`f_\chi ^{Ca}`$ obtained respectively by projecting the interaction $`\mathrm{\Gamma }_C\mathrm{\Lambda }_a`$ onto chiral and diquark channels,
$`A=2\left({\displaystyle \underset{Ca}{}}\mathrm{\Sigma }_{Ca}^2f_\mathrm{\Delta }^{Ca}\right)^1,B=2\left({\displaystyle \underset{Ca}{}}\mathrm{\Sigma }_{Ca}^2f_\chi ^{Ca}\right)^1.`$ (24)
To make contact to microscopic theories, we note that a small coupling limit in either channel corresponds to a small Fierz constant and hence to large parameters $`A`$ or $`B`$. This limit favors small fields $`\mathrm{\Delta }`$ or $`\sigma `$. Because we can always rescale the condensation fields by either $`\sqrt{A}`$ or $`\sqrt{B}`$, the only independent parameter in Eq. (23) is the ratio of $`B/A`$, which by virtue of Eq. (24) is a measure of the balance between the condensation forces. Again, Hermitian matrices $``$ satisfy $`B/AN_c/2`$.
The mass $`m`$ and the parameter $`\eta `$ explicitly break chiral and color symmetries. They act as external fields which select a particular direction for the condensation pattern, and should be taken to zero at the end of the calculations. (They can also be kept constant to study the effect of a small external field, a point which we take in Sec. IV.) They are useful for obtaining the order parameters from derivatives of the partition function. In the thermodynamic limit $`N\mathrm{}`$, $`Z(\mu ,T)`$ in Eq. (20) obeys
$`\underset{N\mathrm{}}{lim}\mathrm{log}Z(\mu ,T)=\underset{N\mathrm{}}{lim}4N\underset{\sigma ,\mathrm{\Delta }}{\mathrm{min}}\left\{\mathrm{\Omega }(\sigma ,\mathrm{\Delta })\right\},`$ (25)
where the right side represents the global minimum of $`\mathrm{\Omega }`$ in Eq. (23) for fixed $`\mu `$ and $`T`$. The order parameters for chiral and diquark condensations are given as
$`\psi ^{}\psi =\underset{m0}{lim}\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{4N_fN}}{\displaystyle \frac{\mathrm{log}Z}{m}}|_{\eta ,\eta ^{}=0},\psi ^T\psi =\underset{\eta ,\eta ^{}0}{lim}\underset{N\mathrm{}}{lim}{\displaystyle \frac{1}{4N}}{\displaystyle \frac{\mathrm{log}Z}{\eta ^{}}}|_{m=0},`$ (26)
where the number of flavor is $`N_f=2`$. Note that the thermodynamic limit $`N\mathrm{}`$ must be taken first before the small field limit $`m,\eta 0`$, see . Given the $`m`$\- and $`\eta `$-dependences of the log terms in Eq. (23), we have
$`\psi ^{}\psi =B\sigma (\mu ,T),\psi ^T\psi =A\mathrm{\Delta }(\mu ,T),`$ (27)
where $`\sigma (\mu ,T)`$ and $`\mathrm{\Delta }(\mu ,T)`$ are the condensation fields which minimize $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ for fixed $`\mu `$ and $`T`$.
Equation (23) is the main expression from which we will deduce the phase structure as a function of the ratio $`B/A`$. Its form is very simple to understand. The quadratic terms correspond to the energy cost for creating static field configurations with finite $`\sigma `$ and $`\mathrm{\Delta }`$. The log terms represent the energy of interaction between the condensation fields and the quark degrees of freedom. They can be written in a compact form as $`\mathrm{Tr}[\mathrm{log}S(\sigma ,\mathrm{\Delta })]=\mathrm{log}\mathrm{det}S(\sigma ,\mathrm{\Delta })`$, where $`\mathrm{Tr}`$ is a trace in flavor, spin, and color, and $`S`$ is the single quark propagator in a background of $`\sigma `$ and $`\mathrm{\Delta }`$ fields. Substituting the Mastubara frequency $`Tip_4`$, it becomes clear that the poles of $`S`$ in $`p_4`$ correspond to the excitation energies of the system. From Eq. (23), we see that two colors develop gapped excitations with
$`p_4`$ $`=`$ $`\pm \sqrt{(\sigma \mu )^2+\mathrm{\Delta }^2},`$ (28)
where the plus and minus signs respectively correspond to particle and antiparticle modes. The $`N_c2`$ remaining colors have ungapped excitation with
$`p_4`$ $`=`$ $`\pm |\sigma \mu |.`$ (29)
It is important to recognize that the potential $`\mathrm{\Omega }`$ in Eq. (23) contains the contribution to the thermodynamics from only the low energy modes of Eqs. (28) and (29). The right side of Eq. (23) thus corresponds to the non-analytic piece in the thermodynamic potential which describes the critical physics related to chiral and color symmetry breaking. In a microscopic model, the right side of Eq. (8) would also contain a smooth analytic component $`\mathrm{\Omega }_{\mathrm{reg}}(\mu ,T)`$ which arises from all other, non-critical degrees of freedom in the system. Our model does not contain this contribution and thus cannot be taken as a quantitative description of bulk thermodynamics properties; the model is constructed to describe the critical properties and should be used as such.
We show in the next section that the forms of the potential $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ in Eq. (23) and of the associated excitation energies in Eqs. (28) and (29) are sufficient to produce a rich variety of phase diagrams which illustrate in a clear way the interplay between chiral and color symmetries. In particular, we will find that single-gluon exchange reproduces the topology obtained in many microscopic models of finite density QCD, see for instance .
## III Exploring the phase diagrams
We noted earlier that the potential $`\mathrm{\Omega }`$ in Eq. (23) is very similar to a Landau-Ginzburg functional. Although $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ is not an algebraic function, it is equivalent to a polynomial of order $`\sigma ^6`$ and $`\mathrm{\Delta }^4`$: The gap equations reduce to coupled polynomial equations of fifth order in $`\sigma `$ and third order in $`\mathrm{\Delta }`$. Generically, four types of solutions exist:
* (i) the $`0`$-phase, the trivial phase in which both $`\sigma `$ and $`\mathrm{\Delta }`$ vanish,
* (ii) the $`\chi `$-phase, in which chiral symmetry is spontaneously broken but $`\mathrm{\Delta }=0`$,
* (iii) the $`\mathrm{\Delta }`$-phase, in which color symmetry is spontaneously broken but $`\sigma =0`$,
* (iv) the $`\chi \mathrm{\Delta }`$ phase, a mixed broken symmetry phase in which both fields are non-vanishing, $`\sigma 0`$ and $`\mathrm{\Delta }0`$.
The $`\chi \mathrm{\Delta }`$-phase is thermodynamically distinct from the $`\chi `$\- and $`\mathrm{\Delta }`$-phases and is not a โmixtureโ of these two phases.
At a given $`\mu `$, $`T`$, and $`B/A`$, each of the solutions (i) - (iv) can either be a minimum or a saddle-point of $`\mathrm{\Omega }`$. The complex flow of these solutions with the variation of $`B/A`$ leads to large variety of phase diagrams. The phase structures can however be grouped according to their topologies as shown in Figs. 16We have rescaled in each figure the units of $`T`$ and $`\mu `$ for clarity. Figure 1 shows the case of smallest values $`B/A`$, which favor chiral over diquark condensation. $`B/A`$ then increases continuously from Fig. 2 to 6. We now discuss Figs. 1-6 in the six following subsections. We indicate in parentheses the corresponding ranges of $`B/A`$ and discuss their limits in the text.
Pure chiral condensation: ($`0B/A0.139N_c`$ or $`0B/A0.418`$ for $`N_c=3`$; see Fig. 1). Generically, the global minimum of $`\mathrm{\Omega }`$ in Eq. (23) is realized by field configurations which maximize the log terms while keeping reasonably low values of the quadratic terms $`A\mathrm{\Delta }^2+B\sigma ^2`$. We consider first the limit $`AB`$. The minimum of $`\mathrm{\Omega }`$ must then correspond to $`\mathrm{\Delta }=0`$ in order to avoid the large energy penalty $`A\mathrm{\Delta }^2`$. No diquark condensation occurs in this case, and color enters only as a prefactor $`N_c`$ in the number of degrees of freedom. This becomes clear if we absorb $`N_c`$ into $`B`$ and set $`B`$ to $`1`$; we then recover the potential $`\mathrm{\Omega }`$ studied in the chiral random matrix models which neglect color altogether.
We briefly recall the phase structure in this case. The gap equation for the chiral field, $`\mathrm{\Omega }(\sigma ,0)/\sigma =0`$, has a trivial root $`\sigma =0`$ and four other roots which satisfy the following quadratic equation for $`\sigma ^2`$
$`N_c(\mu ^2T^2)+B(\mu ^2+T^2)^2+(2B(T^2\mu ^2)N_c)\sigma ^2`$ $`+`$ $`B\sigma ^4=0.`$ (30)
In the high temperature limit $`T\mu `$, both roots in $`\sigma ^2`$ are negative. The only real solution of the gap equation is $`\sigma =0`$, and the system is in the symmetric phase. Decreasing $`T`$ for fixed $`\mu `$, one encounters a line of second-order phase transitions
$`L_{\chi ,2}N_c(\mu ^2T^2)+B(\mu ^2+T^2)^2`$ $`=`$ $`0,`$ (31)
where one of the roots $`\sigma ^2`$ in Eq. (30) vanishes. Below $`L_{\chi ,2}`$, the trivial root becomes a local maximum of $`\mathrm{\Omega }(\sigma ,0)`$, and we have a pair of local minima at the real roots $`\sigma =\pm \sigma _0`$, where
$`\sigma _0=\left({\displaystyle \frac{N_c}{2B}}T^2+\mu ^2+{\displaystyle \frac{\sqrt{N_c^216B^2\mu ^2T^2}}{2B}}\right)^{1/2}.`$ (32)
These roots correspond to a chiral broken phase. They become degenerate with $`\sigma =0`$ on the second-order line $`L_{\chi ,2}`$, where the potential $`\mathrm{\Omega }`$ scales as $`\mathrm{\Omega }(\sigma ,0)\mathrm{\Omega }(0,0)\sigma ^4`$. Thus, the critical exponents near $`L_{\chi ,2}`$ are those of a mean-field $`\varphi ^4`$ theory.
The second-order line ends at a tricritical point $`(\mu _3,T_3)`$ at which all five roots of the gap equation vanish and where $`\mathrm{\Omega }(\sigma ,0)\mathrm{\Omega }(0,0)\sigma ^6`$, giving now critical exponents of a mean-field $`\varphi ^6`$ theory. From Eq. (30), this happens when $`2B(T_3^2\mu _3^2)N_c=0`$, which with the use of Eq. (31) gives
$`\mu _3=\sqrt{{\displaystyle \frac{N_c}{4B}}}\sqrt{\sqrt{2}1},T_3=\sqrt{{\displaystyle \frac{N_c}{4B}}}\sqrt{\sqrt{2}+1}.`$ (33)
For $`\mu >\mu _3`$, the transition between the chiral and trivial phases is first-order and takes place along the line of equal pressure
$`L_{\chi ,1}`$ $`{\displaystyle \frac{N_c}{2}}\left(1+\sqrt{116\mu ^2T^2{\displaystyle \frac{B^2}{N_c^2}}}\mathrm{log}\left[{\displaystyle \frac{N_c^2}{2B^2}}\left(1+\sqrt{116\mu ^2T^2{\displaystyle \frac{B^2}{N_c^2}}}\right)\right]\right)`$ (35)
$`+N_c\mathrm{log}\left[\mu ^2+T^2\right]+B(\mu ^2T^2)=0.`$
This line intercepts the $`T=0`$ axis at $`\mu =\mu _1`$ which obeys
$`1+{\displaystyle \frac{B\mu _1^2}{N_c}}+\mathrm{log}({\displaystyle \frac{B\mu _1^2}{N_c}})=0.`$ (36)
This gives $`\mu _1=0.528\sqrt{N_c/B}`$, or $`\mu _1=0.914/\sqrt{B}`$ for $`N_c=3`$.
$`L_{\chi ,1}`$ in Eq. (35) is a triple line. To see this and clarify the character of $`(\mu _3,T_3)`$, it is useful to consider the effect of a non-zero quark current mass $`m`$ . A mass $`m0`$ selects a particular direction for chiral condensation. If we now consider the three-dimensional parameter space $`(\mu ,T,m)`$, the region delimited by $`L_{\chi ,2}`$ and $`L_{\chi ,1}`$ in the plane $`m=0`$ thus appears to be a surface of coexistence of the two ordered phases with the chiral fields $`\pm \sigma _0`$ of Eq. (32). Along $`L_{\chi ,1}`$, this surface meets two other โwingโ surfaces which extend symmetrically into the regions $`m>0`$ and $`m<0`$. Each of the wings is a coexistence surface between one of the ordered phases whose chiral field continues to $`\sigma =\pm \sigma _0`$ as $`m0`$ and the high temperature phase. Hence, $`L_{\chi ,1}`$ marks the coexistence of three phases and is a triple line. The three phases become identical at $`(\mu _3,T_3)`$, which is thus a tricritical point. The second- and first-order lines $`L_{\chi ,1}`$ and $`L_{\chi ,2}`$ join tangentially at $`(\mu _3,T_3)`$, see .
The onset of diquark condensation modifies the topology that we have just described. This takes place for coupling ratios $`B/A0.139N_c`$ to which we now turn.
The QCD case: ($`0.139N_cB/A\alpha _1(N_c)`$ or $`0.418B/A1.05`$ for $`N_c=3`$; see Fig. 2). To understand the conditions for the onset of diquark condensation, consider a pure diquark phase by setting $`\sigma =0`$ in Eq. (23). The gap equation $`\mathrm{\Omega }/\mathrm{\Delta }=0`$ then has three solutions: $`\mathrm{\Delta }=0`$, and $`\mathrm{\Delta }=\pm \mathrm{\Delta }_0`$ where $`\mathrm{\Delta }_0=\sqrt{2/A\mu ^2T^2}`$. In the high $`T`$ and $`\mu `$ phase, only the trivial root is real and the system is in the symmetric phase. Inside the semicircle
$`L_{\mathrm{\Delta },2}2/A\mu ^2T^2=0,`$ (37)
the trivial root becomes a maximum of $`\mathrm{\Omega }(0,\mathrm{\Delta })`$, and we have a pair of two real minima $`\mathrm{\Delta }=\pm \mathrm{\Delta }_0`$ for which color symmetry is spontaneously broken. The roots $`\pm \mathrm{\Delta }_0`$ go continuously to zero as one approaches $`L_{\mathrm{\Delta },2}`$, which is thus a second-order line provided no other phase develops. The fate of other ordering forms depends on the pressure in the diquark phase,
$`P_\mathrm{\Delta }=\mathrm{\Omega }_\mathrm{\Delta }`$ $`=`$ $`A(\mu ^2+T^2)2+(N_c2)\mathrm{log}\left(\mu ^2+T^2\right)+2\mathrm{log}\left({\displaystyle \frac{2}{A}}\right).`$ (38)
The maximum pressure $`P_\mathrm{\Delta }`$ is reached on $`L_{\mathrm{\Delta },2}`$, where it also equals the pressure of the trivial phase. The condition for the onset of diquark condensation is then clear: The semicircle must lie in part outside the region occupied by the chiral phase. Then, the maximum pressure in the diquark phase is higher than that of the chiral phase, and the diquark phase is stable near the semicircle.
To find the minimum ratio $`B/A`$ for which this condition holds, we compare the dimensions of the chiral phase to the radius of the semicircle, $`\mu _{\mathrm{semi}}\sqrt{2/A}`$. When $`B/A1`$, the line $`L_{\mathrm{\Delta },2}`$ lies well inside the boundaries of the chiral phase, whose linear dimensions are $`T=\sqrt{N_c/B}`$ (along $`\mu =0`$) and $`\mu _1=0.528\sqrt{N_c/B}`$ (along $`T=0`$). The onset of diquark condensation requires therefore $`B/A`$ to be large enough that the semicircle crosses the first-order line between chiral and trivial phases. This takes place along $`T=0`$ when $`\mu _{\mathrm{semi}}=\mu _1`$, or
$`\sqrt{2/A}=0.528\sqrt{N_c/B},`$ (39)
which gives $`B/A=0.139N_c`$, or $`B/A=0.418`$ for $`N_c=3`$.
For larger values of $`B/A`$, the diquark phase exists in a region delimited by $`L_{\mathrm{\Delta },2}`$, on which it coexists with the symmetric phase, and by a first-order line of equal pressure with the chiral phase. This phase diagram is realized for ratios including that corresponding to single-gluon exchange, $`B/A=N_c/(2(N_c1))`$ (or $`3/4`$ for $`N_c=3`$), which is the ratio taken in Fig. 2. We have verified that the two segments of first-order lines, between the chiral and the diquark phases on the one hand and the chiral and trivial phases on the other hand, join tangentially. Figure 7 shows the chiral and diquark fields as a function of $`\mu `$ and $`T`$. It is worth noting that the chiral field vanishes with a square root law near the second-order line $`L_{\chi ,2}`$ and is discontinuous along the first-order line $`L_{\chi ,1}`$. The diquark field vanishes with a square root law all along the semicircle line.
The topology in Fig. 2 can also be summarized by a simple counting argument. Consider the difference between the pressures in the diquark and the chiral phases along the $`T=0`$ axis,
$`\mathrm{\Delta }\mathrm{\Omega }=\mathrm{\Omega }_\chi \mathrm{\Omega }_\mathrm{\Delta }=N_c2+(A+B)\mu ^22\mathrm{log}{\displaystyle \frac{A\mu ^2}{2}}+N_c\mathrm{log}{\displaystyle \frac{B\mu ^2}{N_c}}.`$ (40)
This difference varies from $`\mathrm{\Delta }\mathrm{\Omega }=N_c(1+(2B)/(N_cA)+\mathrm{log}[(2B)/(N_cA)])`$ as $`\mu \sqrt{2/A}`$, to $`\mathrm{\Delta }\mathrm{\Omega }=(N_c2)\mathrm{log}\mu ^2`$ as $`\mu 0`$. We have just argued that $`\mathrm{\Delta }\mathrm{\Omega }`$ can be negative in the former limit if $`B/A>0.139N_c`$. In the latter limit, however, $`\mathrm{\Delta }\mathrm{\Omega }`$ is always negative for $`N_c>2`$ and the system is necessarily in the chiral phase. That the diquark phase must be metastable for small $`\mu `$ is clearly a consequence of the fact that chiral condensation uses all $`N_c`$ colors while diquark condensation uses only colors $`1`$ and $`2`$. This counting argument was mentioned in early models of color superconductivity and is here completely manifest. For $`N_c>2`$, the diquark phase must appear at densities higher than those appropriate for the chiral phase.
Onset of the mixed broken symmetry phase: ($`\alpha _1(N_c)B/AN_c/\sqrt{8}`$ or $`1.05B/A1.06`$ for $`N_c=3`$; see Fig. 3). The chiral solution $`\sigma _0`$, Eq. (32), ceases to be a local minimum of $`\mathrm{\Omega }(\sigma ,0)`$ for sufficiently large $`\mu `$. It turns into a saddle point along a line where the second derivative $`^2\mathrm{\Omega }/\mathrm{\Delta }^2`$ vanishes,
$`L_{\mathrm{mix}}\left(N_cA2B\right)\left(N_c+\sqrt{N_c^216B^2\mu ^2T^2}\right)8B^2\mu ^2=0.`$ (41)
A new minimum of $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ with both $`\sigma 0`$ and $`\mathrm{\Delta }0`$ develops in the region to the right of $`L_{\mathrm{mix}}`$. This new minimum corresponds to a new phase, the $`\chi \mathrm{\Delta }`$-phase, which competes with the diquark phase. When the $`\chi \mathrm{\Delta }`$-phase first appears along $`L_{\mathrm{mix}}`$, it has the same pressure as the chiral phase. Therefore, the condition for this new phase to realize the largest pressure is that the instability line $`L_{\mathrm{mix}}`$ lies in the region spanned by the chiral phase. Then, along $`L_{\mathrm{mix}}`$, the new phase necessarily has a pressure that exceeds that in the diquark phase and is favored.
To see what ratios $`B/A`$ are needed for the mixed broken symmetry phase, consider the point where $`L_{\mathrm{mix}}`$ meets the $`T=0`$ axis,
$`\mu _{\mathrm{mix}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{N_c}{2B}}\left({\displaystyle \frac{N_cA}{2B}}1\right)}.`$ (42)
When $`B/A1`$, $`\mu _{\mathrm{mix}}\sqrt{N_c/B}/\sqrt{4B/A}\sqrt{N_c/B}`$ and $`L_{\mathrm{mix}}`$ lies well outside the chiral phase. $`L_{\mathrm{mix}}`$ then moves towards the $`\chi `$-phase as $`B/A`$ increases. It first crosses the first-order line between chiral and diquark phases when $`\mu _{\mathrm{mix}}=\mu _1`$, where $`\mu _1`$ is the point of equal pressure which obeys $`\mathrm{\Delta }\mathrm{\Omega }=0`$ in Eq. (40). This condition gives a ratio $`B/A=\alpha _1(N_c)`$, where $`\alpha _1`$ is a non-trivial function of $`N_c`$ which we study in the appendix. Here, we note only that $`\alpha _1(3)=1.05`$ for three colors. When $`B/A>\alpha _1(N_c)`$, we find that the mixed broken symmetry phase develops in the wedge bordered by the second-order line $`L_{\mathrm{mix}}`$ of Eq. (41), and a first-order line on which it coexists with the diquark phase. The phase diagram is shown in Fig. 3, while an expanded view of the wedge of mixed broken symmetry and of the first-order line near $`(\mu _3,T_3)`$ are respectively shown in Figs. 8 and 9.
A new critical point: ($`N_c/\sqrt{8}B/A\alpha _2(N_c)`$ or $`1.06B/A1.163`$ for $`N_c=3`$; see Fig. 4). As $`B/A`$ increases above $`\alpha _1(N_c)`$, the semicircle $`2/A=\mu ^2+T^2`$ grows relative to the chiral phase, and eventually reaches the tricritical point $`(\mu _3,T_3)`$ of Eq. (33) when $`B/A=N_c/\sqrt{8}`$. At this stage, the segment of first-order line between chiral and trivial phases disappears. For $`B/A>N_c/\sqrt{8}`$, the semicircle meets the second-order line between chiral and trivial phases, $`L_{\chi ,2}`$ in Eq. (31), at a new critical point
$`\mu _4=\sqrt{{\displaystyle \frac{1}{A}}\left(1{\displaystyle \frac{2B}{N_cA}}\right)},T_4=\sqrt{{\displaystyle \frac{1}{A}}\left(1+{\displaystyle \frac{2B}{N_cA}}\right)},`$ (43)
which now separates three phases as shown in Fig. 4.
Coexistence of four phases: ($`\alpha _2(N_c)B/AN_c/2`$ or $`1.163B/A1.5`$ for $`N_c=3`$; see Fig. 5). With still higher coupling ratios $`B/A`$, the wedge of mixed broken symmetry in Fig. 4 grows in size relative to the other phases. Its tip reaches the new critical point $`(\mu _4,T_4)`$ when $`B/A`$ satisfies the condition
$`{\displaystyle \frac{B}{A}}=\alpha _2(N_c){\displaystyle \frac{4N_cN_c^{3/2}\sqrt{2N_c4}}{4(4N_c)}},`$ (44)
the derivation of which we detail in the appendix. We just note here that $`\alpha _2(3)=1.163`$. This coupling ratio also marks the appearance of two new critical points, as illustrated in Fig. 5. First, when $`B/A>\alpha _2(N_c)`$, the point $`(\mu _4,T_4)`$ characterizes the coexistence of all four phases, and it has thus become a tetracritical point. The pressure of the system at $`(\mu _4,T_4)`$ has the form $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })\mathrm{\Omega }(0,0)a\sigma ^4+b\mathrm{\Delta }^4+c\sigma ^2\mathrm{\Delta }^2`$, where $`a`$, $`b`$, and $`c`$ are constants detailed in the appendix. We note in particular that none of the four second-order lines in Fig. 5 join tangentially at the tetracritical point.
The second point appearing above $`B/A=\alpha _2(N_c)`$ is a tricritical point $`(\mu _{3\mathrm{m}},T_{3\mathrm{m}})`$ which lies on the boundary between the $`\chi \mathrm{\Delta }`$\- and the $`\mathrm{\Delta }`$-phases. Its origin can be understood from the similarity between the characters of the mixed broken symmetry phase and the chiral phase. The diquark field in the $`\chi \mathrm{\Delta }`$-phase is
$`\mathrm{\Delta }(\sigma )`$ $`=`$ $`\left({\displaystyle \frac{1+\sqrt{1+4A^2\mu ^2\sigma ^2}}{A}}\mu ^2T^2\sigma ^2\right)^{1/2}.`$ (45)
Substituting $`\mathrm{\Delta }`$ in $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ for this expression, we find that the gap equation for the chiral field, $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta }(\sigma ))/\sigma =0`$, has five roots. Despite the fact that $`\mathrm{\Delta }(\sigma )0`$, the dynamics of these roots as a function of $`\mu `$ and $`T`$ is identical to that of the roots in the pure chiral condensation case discussed at the beginning of this section. In particular, the transition to the $`\mathrm{\Delta }`$-phase starts out second-order near $`(\mu _4,T_4)`$ and takes place along a line where three of the five roots vanish. This second-order line is thus determined by the condition
$`{\displaystyle \frac{d^2\mathrm{\Omega }}{d\sigma ^2}}(\sigma ,\mathrm{\Delta }(\sigma ))|_{\sigma 0}`$ $`=`$ $`0,`$ (46)
which is the analog of Eq. (31) and gives
$`L_{\chi \mathrm{\Delta }\mathrm{\Delta },2}(N_c2)(\mu ^2T^2)+\left(BA(1A\mu ^2)\right)\left(\mu ^2+T^2\right)^2=0.`$ (47)
In analogy with the pure chiral case, the transition becomes first order at high $`\mu `$; second- and first-order segments join tangentially at a tricritical point $`(\mu _{3\mathrm{m}},T_{3\mathrm{m}})`$.
We determine the location of this point as follows. We observe that, as in the pure chiral broken case, the potential along the second-order line $`L_{\chi \mathrm{\Delta }\mathrm{\Delta },2}`$ has the form $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta }(\sigma ))\mathrm{\Omega }(0,0)\sigma ^4`$. (See Eq. (46).) This scaling form becomes $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta }(\sigma ))\mathrm{\Omega }(0,0)\sigma ^6`$ at $`(\mu _{3\mathrm{m}},T_{3\mathrm{m}})`$ where all five roots vanish. Thus, the tricritical point is located on the line $`L_{\chi \mathrm{\Delta }\mathrm{\Delta },2}`$ at the point where
$`{\displaystyle \frac{d^4\mathrm{\Omega }}{d\sigma ^4}}(\sigma ,\mathrm{\Delta }(\sigma ))`$ $`=`$ $`0,`$ (48)
which is the analog of the condition $`2B(T^2\mu ^2)N_c=0`$ for the pure chiral case. We have solved Eq. (48) numerically to determine $`(\mu _{3\mathrm{m}},T_{3\mathrm{m}})`$ in Figs. 5 and 6.
Disappearance of the chiral phase. ($`\alpha _2(N_c)B/A`$; see Figs. 6 and 10). The wedge of mixed broken symmetry gains space in the $`(\mu ,T)`$ plane at the expense of the chiral phase as $`B/A`$ increases above $`\alpha _2(N_c)`$. The chiral phase eventually shrinks to a vertical line along the $`\mu =0`$ axis when $`B/A=N_c/2`$, as shown in Fig. 10. That the chiral phase is still present for $`\mu =0`$ can be understood as follows. When $`B/A=N_c/2`$ and $`\mu =0`$, both condensation fields appear in Eq. (23) in the combination $`\sigma ^2+\mathrm{\Delta }^2`$. A pure diquark solution $`(\sigma ,\mathrm{\Delta })=(0,\mathrm{\Delta }_0)`$ can thus always be rotated into a pure chiral solution $`(\sigma ,\mathrm{\Delta })=(\mathrm{\Delta }_0,0)`$ , and thus does not represent an independent phase. This symmetry is however no longer present as soon as $`\mu 0`$, in which case the diquark phase becomes thermodynamically independent from the chiral phase.
Recall that the ratio $`B/A=N_c/2`$ is the maximum ratio that our model can realize . It is instructive however to explore higher coupling ratios, although they do not necessarily describe physical situations, by setting $`B/A>N_c/2`$ by hand in Eq. (23). We find that higher $`B/A`$ force the diquark phase to grow in size at the expense of the mixed broken symmetry phase. Figure 6 shows one example with $`B/A=1.8`$ and $`N_c=3`$. The similarities between the critical properties of the mixed broken symmetry phase and those of the pure chiral broken phase are now clear: Compare for instance the critical lines between Figs. 1 and 6.
## IV Discussion
Baryon density discontinuity. We have argued that the potential $`\mathrm{\Omega }`$ in Eq. (8) represents the non-analytic contribution to the thermodynamics which is directly associated with the breaking of chiral and color symmetry. The potential $`\mathrm{\Omega }`$ should therefore not be used too literally in computations of the bulk properties of the phases we have encountered. It may however give reasonable estimates of discontinuities near a phase transition. For instance in the pure chiral case, the random matrix models of Ref. estimate that the baryon density $`n_B=(1/3)\mathrm{\Omega }(\mu ,T)/\mu `$ changes discontinuously at the first order point along $`T=0`$ by an amount $`\mathrm{\Delta }n_B2.5n_0`$. Here, $`n_0=0.17\mathrm{fm}^3`$ is the density of normal nuclear matter. The result for $`\mathrm{\Delta }n_B`$ relies on an evaluation of the number of degrees of freedom $`N`$ from instanton models and seems to be a reasonable estimate of the baryon density discontinuity .
$`\mathrm{\Delta }n_B`$ is modified by the presence of diquark condensation. We consider single-gluon exchange with $`N_c=3`$, which realizes $`B/A=3/4`$, and work in the limit $`T=0`$. Taking the derivative of Eq. (40) with respect to $`\mu `$, we find the discontinuity in baryon density at the point $`\mu _1`$ of equal pressure between chiral and color phases to be
$`N_\mathrm{\Delta }N_\chi `$ $`=`$ $`{\displaystyle \frac{2}{\mu _1}}\left(1+{\displaystyle \frac{7}{3}}B\mu _1^2\right).`$ (49)
Here, $`\mu _1`$ obeys $`\mathrm{\Omega }_\chi =\mathrm{\Omega }_\mathrm{\Delta }`$ in Eq. (40); we have $`\mu _1=0.87/\sqrt{B}`$. Hence, $`N_\mathrm{\Delta }N_\chi 6.4\sqrt{B}`$ in the appropriate unit of inverse chemical potential. Were no diquark condensation to occur, we would have found a discontinuity
$`N_0N_\chi `$ $`=`$ $`{\displaystyle \frac{6}{\mu _1}}\left(1+{\displaystyle \frac{B\mu _1^2}{3}}\right),`$ (50)
where $`\mu _1`$ is now the point of equal pressure between chiral and trivial phases. We want to keep the same $`B`$ as in Eq. (49) so as to compare two situations which realize the same vacuum chiral field $`\sigma _0=\sqrt{3/B}`$. The condition of equal pressure in Eq. (36) gives then $`\mu _1=0.914/\sqrt{B}`$ and we have $`N_0N_\chi 8.4\sqrt{B}`$. Thus, we find that diquark condensation reduces the discontinuity in baryon density by roughly twenty five percent.
Away from the chiral limit. We now turn to study the effects of a small quark current mass $`m`$ in Eq. (23). For $`m0`$, chiral symmetry is explicitely broken, and the chiral condensate $`\psi ^{}\psi `$ ceases to be a good order parameter. This affects the phase diagrams in Figs. 1-6 in a number of ways. We consider the effect of a small mass $`m`$ chosen so that $`m10`$ MeV in units for which the vacuum chiral field is $`\sigma \sqrt{3/B}400`$ MeV and illustrate a few cases in Figs. 11-14.
Figure 11 shows the limit of small ratios $`B/A`$ which favor chiral over diquark condensation. Since $`\psi ^{}\psi `$ is no longer a good order parameter, any two given points in the phase diagram can be connected by a trajectory along which no thermodynamic discontinuity occurs. It results that the second-order line $`L_{\chi ,2}`$ in Eq. (31) is no longer present when $`m0`$. There remains, however, a first-order line, which ends at a regular critical point $`(\mu _c,T_c)`$. This point can be located as follows. Along the first-order line, the potential $`\mathrm{\Omega }(\sigma ,0)`$ has two minima of equal depth separated by a single maximum. All three extrema become degenerate at the critical point $`(\mu _c,T_c)`$, past which $`\mathrm{\Omega }(\sigma ,0)`$ possesses only one minimum. The location of $`(\mu _c,T_c)`$ can thus be determined from the condition that $`\mathrm{\Omega }(\sigma ,0)`$ scales as
$`\mathrm{\Omega }(\sigma ,0)\mathrm{\Omega }(0,0)(\sigma \sigma _0)^4,`$ (51)
at $`(\mu _c,T_c)`$ and for small deviations $`|\sigma \sigma _0|`$. The location of $`(\mu _c,T_c)`$, as well as $`\sigma _0`$, are then determined by requiring the first three derivatives of $`\mathrm{\Omega }(\sigma ,0)`$ to vanish at the critical point and for $`\sigma =\sigma _0`$. A small mass tends to increase the pressure in the low density โphaseโ with respect to that in the high density โphaseโ. Its results that a mass $`m`$ displaces the first-order line $`L_{\chi ,1}`$ of Eq. (35) to higher $`\mu `$ by an amount linear in $`m`$. The pressure increase also delays the onset of diquark condensation; the ratio $`B/A`$ for which the diquark appears first increases linearly with $`m`$.
Figure 12 shows the case of QCD: there is now a second-order line $`L_{\mathrm{\Delta },2}`$ which separates the high temperature phase from a mixed broken symmetry phase; the dominant effect of $`m`$ on the diquark phase in Fig. 2 is to produce a small chiral field $`\sigma m`$. The effect on the thermodynamics in both the chiral and the trivial phases is, however, second order in $`m`$, and so is the displacement of the second-order line $`L_{\mathrm{\Delta },2}`$ from Fig. 2 to Fig. 12.
The evolution of the phase diagram for higher $`B/A`$ parallels the evolution we have outlined for $`m=0`$ in Figs. 3-6. The phase with a finite diquark field grows until it eventually reaches the critical point $`(\mu _c,T_c)`$. Higher ratios lead to a phase structure in which the first- and second-order lines intersect as shown in Fig. 13. A wedge of another mixed broken symmetry phase, initially with a large chiral field and a small diquark field, appears at still higher $`B/A`$ on the left of the first-order line. This wedge grows in size until it encounters the boundary of the other mixed broken symmetry phase. The second-order lines then merge into a single continuous line as shown in Fig. 14. This line is the locus of point for which
$`{\displaystyle \frac{^2\mathrm{\Omega }(\sigma ,\mathrm{\Delta })}{\mathrm{\Delta }^2}}|_{\mathrm{\Delta }=0}=0,`$ (52)
and on which a chiral solution with a vanishing diquark field turns into a saddle-point of $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$. Inside this boundary, the global minimum of $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ describes a single mixed broken symmetry phase. This phase again exhibits properties similar to those of a chiral phase. It contains in particular a first-order line which ends at a critical point $`(\mu _{\mathrm{cm}},T_{\mathrm{cm}})`$ at which the potential $`\mathrm{\Omega }`$ scales as $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta }(\sigma ))\mathrm{\Omega }(0,0)(\sigma \sigma _0)^4`$, where $`\mathrm{\Delta }(\sigma )`$ is the diquark field for fixed $`\sigma `$.
To summarize, the effects of a small mass is linear for the chiral and mixed broken symmetry phases and quadratic for the diquark and trivial phases. The chiral field no longer represents a good order parameter, and the second-order lines of vanishing second derivatives with respect to $`\sigma `$, i.e. $`L_{\chi ,2}`$ in Eq. (31) and $`L_{\mathrm{mix}}`$ in Eq. (41), disappear. The tricritical points become regular critical points while the tetracritical point in Fig. 5 disappears.
The $`N_c=2`$ and $`N_c\mathrm{}`$ limits. In order to make connection with known results and to obtain some insight on the dependence of the phase structure on the number of colors, it is interesting to consider the limits $`N_c=2`$ and $`N_c\mathrm{}`$. First of all when $`N_c=2`$, the Hermitian matrix models realize only a single coupling ratio $`B/A=1`$ . We focus on single-gluon exchange for which the potential in Eq. (23) becomes
$`\mathrm{\Omega }_2(\sigma ,\mathrm{\Delta })`$ $`=`$ $`A(\sigma ^2+\mathrm{\Delta }^2)\mathrm{log}[(\sigma \mu )^2+\mathrm{\Delta }^2+T^2]\mathrm{log}[(\sigma +\mu )^2+\mathrm{\Delta }^2+T^2].`$ (53)
A rotational symmetry appears at $`\mu =0`$ as $`\mathrm{\Omega }_2`$ depends on chiral and diquark fields via the combination $`\sigma ^2+\mathrm{\Delta }^2`$. Diquark and chiral phases do not in this case represent independent states. We find for the combined fields that
$`\sigma ^2+\mathrm{\Delta }^2`$ $`=`$ $`{\displaystyle \frac{2}{A}}T^2,`$ (54)
below a critical temperature $`T_c=\sqrt{2/A}`$, while symmetry is restored above $`T_c`$. This rotational symmetry is a consequence of the pseudo-reality of $`SU(2)`$-QCD, a property by which the Dirac operator $`D=i_{\mu a}\gamma ^\mu \lambda ^aA_{\mu a}+m`$ commutes with $`\tau _2C\gamma ^5K`$ where $`\tau _2`$ is the antisymmetric $`2\times 2`$ color matrix and $`K`$ the complex conjugation operator. For $`\mu =0`$, this property permits one to arrange color and flavor symmetries into a higher $`SU(4)`$ symmetry, see for instance .
At finite $`\mu `$, however, the $`SU(4)`$ symmetry is explicitly broken. The global minimum of $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ always has $`\sigma =0`$, and the system prefers diquark condensation over chiral symmetry breaking. This results agrees with instanton models , and lattice calculations . We now have a second-order phase transition from an ordered state with $`\mathrm{\Delta }^2=2/A\mu ^2T^2`$ in the low $`T`$ and $`\mu `$ region to a symmetric phase at high $`T`$ and $`\mu `$.
The opposite limit $`N_c\mathrm{}`$ is more subtle. In microscopic models, it is expected on general grounds that the quark-quark interaction is suppressed with respect to the $`\overline{q}q`$ channel by powers of $`1/N_c`$ . In the present model, we observe that diquark condensation disappears as $`N_c\mathrm{}`$ if the interaction is single-gluon exchange. Its coupling ratio $`B/A=N_c/(2N_c2)1/2`$ as $`N_c\mathrm{}`$. Thus, we have $`B/A0.139N_c`$ and the only possible topology for the phase diagram is that in Fig. 1. Hence, no diquark condensate forms. Other Hermitean interactions with $`B/AO(N_c)`$ can however explore the full range of phase diagrams which display diquark condensation. The actual number of possible toplogies is however reduced to five, as $`\alpha _2(N_c)N_c/\sqrt{8}`$ for $`N_c\mathrm{}`$ (see Appendix) and Fig. 4 can no longer be realized.
We summarize the variation with $`N_c`$ of the coupling ratios characterizing a change in topology and of the ratio realized by single-gluon exchange in Table 1.
Comparison with a microscopic model: In contrast to microscopic models, the random matrix interaction does not lead to a logarithmic instability of the gap equation near the Fermi surface. To clarify this effect, we consider diquark condensation in a chiral symmetric phase at $`T=0`$ and compare the random matrix approach to a microscopic model. We choose here the NJL study of Berges and Rajagopal . In the random matrix formulation the gap equation
$`2A\mathrm{\Delta }`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }}{\mathrm{\Delta }^2+\mu ^2}}`$ (55)
has a non-trivial root $`\mathrm{\Delta }\sqrt{\mu _c\mu }`$ with $`\mu _c=\sqrt{2/A}`$. The model of gives an equation of the form
$`2A^{}\mathrm{\Delta }`$ $`=`$ $`\mathrm{\Delta }{\displaystyle _0^{\mathrm{}}}๐qq^2F^4(q)\left({\displaystyle \frac{1}{\sqrt{(q\mu )^2+\mathrm{\Delta }^2F^4(q)}}}+{\displaystyle \frac{1}{\sqrt{(q+\mu )^2+\mathrm{\Delta }^2F^4(q)}}}\right),`$ (56)
where $`F(q)`$ is an appropriate form factor which falls off over $`q\mathrm{\Lambda }O(\mathrm{\Lambda }_{\mathrm{QCD}})`$. Clearly, the right side of Eq. (56) contains a singularity for $`q\mu `$ and $`\mathrm{\Delta }0`$ which is absent in Eq. (55). This singularity has two consequences. First, the behavior of the diquark field for large $`\mu `$ depends sensitively on the form factor. As $`\mathrm{\Delta }0`$, the singularity at $`q\mu `$ in the right side of Eq. (56) gives $`2A^{}\mu ^2F^4(\mu )\mathrm{log}\mathrm{\Lambda }\mu /\mathrm{\Delta }^2`$ to logarithmic order. Thus, instead of the square root behavior $`\mathrm{\Delta }(\mu )(\mu _c\mu )^{1/2}`$ of the random matrix approach, $`\mathrm{\Delta }`$ now vanishes as $`\mu \mathrm{}`$ as $`\mathrm{\Delta }(\mu )\mathrm{exp}(c/(\mu ^2F^4(\mu )))`$, where $`c`$ is a constant. This result is only true for a smooth cutoff $`F(q)`$. For a sharp cutoff $`F(q)=\mathrm{\Theta }(\mathrm{\Lambda }\mu )`$, where $`\mathrm{\Theta }(x)`$ is the Heavyside function, the diquark field exhibits the square root singularity. This tail is exponentially sensitive to form factors; with regards to establishing the general topology of the phase diagram, such behavior is not qualitatively different from $`\mathrm{\Delta }=0`$. The second consequence of the logarithmic singularity in Eq. (56) is that $`\mathrm{\Delta }(\mu )`$ must be non-monotonic for intermediate $`\mu `$. This profile can be seen by drawing in the plane $`(\mathrm{\Delta },\mu )`$ the lines of constant height for the right side of Eq. (56): these lines must go around the point $`(\mu _0,0)`$ where the logarithmic singularity is the strongest as $`\mathrm{\Delta }0`$, and $`\mathrm{\Delta }(\mu )`$ reaches a maximum at $`\mu _0`$.
Color dependency in the chiral fields. The patterns of symmetry breaking which we have described can become even richer if we allow the chiral fields to depend on color. This possibility arises in the instanton model of Carter and Diakonov , who remarked that the Dyson-Gorkov equations close in color space on the condition that gapped and ungapped quarks can develop different masses. We studied the effects of this additional degree of freedom in the limit of zero chemical potential and found no change in the phase structure for ratios $`B/AN_c/2`$. The situation is different for finite $`\mu `$. Choosing different masses for the two gapped and the $`N_c2`$ ungapped quarks leads in some limits to an increase in the pressure of phases with finite diquark fields. The main consequence is an increase of both the parameter range and the region in the $`(\mu ,T)`$ plane for which the mixed broken symmetry phase exists. This result is obvious since the mixed broken symmetry phase is the only phase which can exploit this additional degree of freedom. However, there is also a general mechanism at work by which a change of phase structure always takes place above a certain threshold for the coupling constants.
We wish to illustrate these effects in a few cases and concentrate on the limit $`N_c=3`$ and $`m=0`$ for clarity. We denote the chiral fields for color $`1`$ and $`2`$ by $`\sigma _1`$ and that for color $`3`$ by $`\sigma _3`$. In order to permit these fields to be different, we include the projection of the interaction onto a chiral-$`\lambda _8`$ channel as described in . For $`\mu 0`$, the thermodynamical potential becomes
$`\mathrm{\Omega }(\sigma _1,\sigma _3,\mathrm{\Delta })`$ $`=`$ $`A\mathrm{\Delta }^2+B(\beta _1\sigma _1^2+\beta _2\sigma _1\sigma _3+\beta _3\sigma _3^2)\mathrm{log}[(\sigma _1\mu )^2+\mathrm{\Delta }^2+T^2]\mathrm{log}[(\sigma _1+\mu )^2+\mathrm{\Delta }^2+T^2]`$ (58)
$`{\displaystyle \frac{1}{2}}\mathrm{log}[(\sigma _3\mu )^2+T^2]{\displaystyle \frac{1}{2}}\mathrm{log}[(\sigma _3+\mu )^2+T^2],`$
where
$`\beta _1={\displaystyle \frac{4}{9}}+{\displaystyle \frac{C}{3B}},\beta _2={\displaystyle \frac{4}{9}}{\displaystyle \frac{2C}{3B}},\beta _3={\displaystyle \frac{1}{9}}+{\displaystyle \frac{C}{3B}},`$ (59)
describe the coupling between $`\sigma _1`$ and $`\sigma _3`$. $`C`$ is a coupling constant associated with the chiral-$`\lambda _8`$ channel . By fine tuning the various Lorentz and color channels which compose the random matrix interactions, we can realize a range of ratios $`B/C`$ for any fixed $`B/A`$<sup>ยง</sup><sup>ยง</sup>ยงThis range is $`3/16B/C3/2`$ for $`B/A<3/4`$ and then reduces linearly to $`B/C=3/2`$ for $`B/A=3/2`$. New patterns of symmetry breaking only appears when $`C>0`$ and the chiral-$`\lambda _8`$ channel is attractive .
Consider first single-gluon exchange. The coupling ratios are $`B/A=3/4`$ and $`C/B=3/16`$. The chiral-$`\lambda _8`$ channel is repulsive, and the phases with the largest pressure always satisfy $`\sigma _1=\sigma _3`$, which is the case shown in Fig. 2. If we now fine tune the interactions so as to increase $`B/C`$ to $`B/C3/2`$ while keeping $`B/A=3/4`$, the phase diagram changes substantially. The first-order line between the chiral and trivail phases splits at the point where it meets the diquark transition line into two first-order lines. Together with the $`T=0`$ axis, they delimitate a wedge of mixed broken symmetry phase with $`\sigma _1\sigma _3`$ and $`\mathrm{\Delta }0`$. Thus, with an attractive channel $`B/C3/2`$ the mixed broken symmetry phase appears much earlier than previously discussed. This is an extreme case. For fixed $`B/A`$, the mixed broken symmetry phase does not appear immediately as $`B/C`$ increases from $`0`$: there is a threshold value above which the new phase appears (this value is $`B/C1.29`$ for $`B/A=0.75`$). Therefore, it takes large variations of the coupling constants $`B/A`$ and $`B/C`$ away from the values expected for single-gluon exchange to modify the phase structure of Fig. 2.
To complete the picture of the effects of an attractive channel $`C>0`$, consider next $`B/A=3/2`$. This ratio can only be realized by a color diagonal interaction for which $`B/C=3/2`$ . In this case, there is no freedom in fine tuning $`B/C`$ to other values. The effect of splitting masses in colors is now maximal. $`\beta _2`$ in Eq. (59) vanishes and $`\sigma _3`$ decouples from the other two fields. Its gap equation leads to the same solutions as a pure chiral phase with $`\sigma 0`$ and $`\mathrm{\Delta }=0`$ and the phase diagram for the third color is that of Fig. 1. For $`B/A=B/C=3/2`$, the partial pressure for colors $`1`$ and $`2`$ has the same form as in the limit $`N_c=2`$, see Eq. (53). There is then a rotational symmetry between $`\sigma _1`$ and $`\mathrm{\Delta }`$ for $`\mu =0`$ and $`\sigma _1`$ vanishes for $`\mu 0`$ while $`\mathrm{\Delta }=2/A\mu ^2T^2`$ for $`\mu ^2+T^22/A`$. The overall phase diagram is obtained by superposing the two pictures; we have a mixed broken symmetry phase with $`\sigma _1=0`$, $`\sigma _30`$, and $`\mathrm{\Delta }0`$ on the left of the first-order line $`L_{\chi ,1}`$ of Eq. (35). This phase is contiguous to a pure diquark phase which develops as in Fig. 2 on the right of $`L_{\chi ,1}`$ and below the semicircle. With respect to Fig. 5, the mixed broken symmetry phase thus occupies a wider area in the $`(\mu ,T)`$ plane.
A color-$`6`$ condensate. It has been suggested that an energy gain may result if the third color condenses in a spin-$`1`$ color symmetric state . We find no such condensation for single-gluon exchange. We can however fine tune the interactions so as to keep $`B/A=3/2`$ and allow this channel to develop. The Fierz constant for projecting on a color $`6`$ is half that for a $`\overline{3}`$ color state. Thus, we can understand qualitatively how a color-$`6`$ behaves by repeating our previous analysis with an additional phase which now may develop inside a semicircle of radius $`\sqrt{2}`$ smaller than the radius of the diquark phase. This semicircle crosses the first-order line beween chiral and diquark phases at $`B/A0.754`$ and a color-$`6`$ phase can, in favorable cases, increase the pressure of the system. However, as far as QCD is concerned, a color-$`6`$ condensate should be very small since its threshold ratio is very close to $`B/A=3/4`$. This is an example of a result which cannot be regarded as robust: in a microscopic theory, the fate of the color-$`6`$ phase will inevitably depend on the details of the interaction and on whether these give rise to exponential tails for the associated condensation field. By contrast, the chiral and diquark phases are fully developed for $`B/A=3/4`$ and the phase diagram in Fig. 2 should thus be considered robust.
## V Conclusions
Our random matrix model leads to a thermodynamic potential which has a very simple form. Yet, it contains enough physics to illustrate in a clear way the interplay between chiral symmetry breaking and the formation of quark Cooper pairs. We have found that this interplay results in a variety of phase diagrams which can be characterized by a total of only six different topologies. Single-gluon exchange leads to the topology shown in Fig. 2, a phase diagram familiar from microscopic models. We have considered the chiral and scalar diquark channels, which seem the most promising ones. We have found that chiral and diquark phases are fully developped for the ratio realized by single-gluon exchange and that it takes large variations in the coupling ratios $`B/A`$ and $`C/B`$ to depart from that result. On the other hand, other less attractive condensation channels seem sensitive to coupling constant ratios and are expected in microscopic models to depend on the details of the interactions. Furthermore, these channels develop weak condensates at best. It is worth keeping in mind that the present picture is mean-field; we expect that quantum fluctuations will inevitably have large effects on the weak channels, which should thus not be considered robust. We conclude that QCD with two light flavors should realize the topology suggested by single-gluon exchange and that this topology is stable against variations in the detailed form of the microscopic interactions.
Many effects lie in the mismatch between the number of colors involved in each order parameter: all $`N_c`$ colors contribute to the chiral field, while a Cooper pair involves two colors. A result of this for $`N_c=3`$ is that the chiral broken phase prevails at low densities. Furthermore, our model reproduces to a reasonable extent the expected limits at small and large $`N_c`$. We expect such counting arguments to be valid in both microscopic models and lattice calculations. More generally, we believe that random matrix models can provide insight into calculations of QCD at finite density by providing simple illustrations for many of the mechanisms which are at work.
Not all the phase diagrams that we have studied are directly relevant to QCD. However, many of their characteristics such as the presence of tricritical and tetracritical points are generic to systems in which two forms of order compete. Our model could naturally be extended to the study of nuclear or condensed matter systems in which such competion takes place. The construction of a genuine theory seems technically involved at first glance but in fact contains only three basic ingredients: the identification of the symmetries at play and their associated order parameters, the knowledge of the elementary excitations in a background of condensed fields, and the calculation of the range of coupling constants realized by the random matrix interactions. These three components are sufficient to build a thermodynamical potential and determine the resulting phase structures in parameter space.
## Acknowledgements
We are grateful for stimulating discussions with G. Carter, D. Diakonov, H. Heiselberg, and K. Splittorff.
## A Calculation of $`\alpha _1(N_c)`$ and $`\alpha _2(N_c)`$
In this appendix, we calculate the ratios $`\alpha _1(N_c)`$ and $`\alpha _2(N_c)`$, which respectively characterize the onset of the $`\chi \mathrm{\Delta }`$-phase and the appearance of the tetracritical point.
### 1 The ratio $`\alpha _1(N_c)`$.
The mixed broken symmetry phase appears first for the coupling ratio $`B/A`$ for which the instability line $`L_{\mathrm{mix}}`$ of Eq. (41) crosses the first-order line between chiral and diquark phases. This crossing occurs on the $`T=0`$ axis. The line $`L_{\mathrm{mix}}`$ meets the $`T=0`$ axis at $`\mu =\mu _{\mathrm{mix}}`$, Eq. (42),
$`\mu _{\mathrm{mix}}^2={\displaystyle \frac{N_c^2A}{4B^2}}{\displaystyle \frac{N_c}{2B}},`$ (A1)
while the condition of equal pressure at that point gives
$`\mathrm{\Omega }_\mathrm{\Delta }\mathrm{\Omega }_\chi =(N_c2)+(A+B)\mu _{\mathrm{mix}}^22\mathrm{log}{\displaystyle \frac{A\mu _{\mathrm{mix}}^2}{2}}+N_c\mathrm{log}{\displaystyle \frac{B\mu _{\mathrm{mix}}^2}{N_c}}=0.`$ (A2)
Combining these two equations gives the determining equation for $`\alpha _1=B/A`$,
$`(N_c2)`$ $`+`$ $`{\displaystyle \frac{N_c}{2\alpha _1^2}}\left({\displaystyle \frac{N_c}{2}}\alpha _1\right)\left(1+\alpha _1\right)2\mathrm{log}\left[{\displaystyle \frac{N_c}{4\alpha _1}}\left({\displaystyle \frac{N_c}{2\alpha _1}}1\right)\right]+N_c\mathrm{log}\left[{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{N_c}{2\alpha _1}}1\right)\right]=0.`$ (A3)
This is a transcendental relation which needs to be solved numerically for each $`N_c`$. In particular, we find $`\alpha _1(2)=1`$, $`\alpha _1(3)=1.05`$, and $`\alpha _1(N_c\mathrm{})0.321N_c`$.
### 2 The ratio $`\alpha _2(N_c)`$.
Two conditions determine $`\alpha _2(N_c)`$. Coming from small $`B/A`$, $`\alpha _2(N_c)`$ corresponds to the ratio at which the wedge of the $`\chi \mathrm{\Delta }`$-phase reaches the critical point $`(\mu _4,T_4)`$ of Eq. (43). Decreasing $`B/A`$ from large values, $`\alpha _2(N_c)`$ marks the coexistence of the tetracritical point $`(\mu _4,T_4)`$ and the tricritical point $`(\mu _{3\mathrm{m}},T_{3\mathrm{m}})`$. We now show that these two conditions are equivalent, and thus that the transition from Fig. 4 to Fig. 5 is continuous.
To proceed, we concentrate on the region near $`(\mu _4,T_4)`$ where we perform a small field expansion of the potential $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ in Eq. (23),
$`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ $``$ $`\mathrm{\Omega }(0,0)+a_0\sigma ^2+b_0\mathrm{\Delta }^2+{\displaystyle \frac{a_1^2}{2}}\sigma ^4+{\displaystyle \frac{b_1^2}{2}}\mathrm{\Delta }^4+c_1\sigma ^2\mathrm{\Delta }^2+๐ช(\mathrm{min}^6(\mathrm{\Delta },\sigma )).`$ (A4)
Here, the coefficients of the quadratic terms are linear in $`\delta \mu =\mu \mu _4`$ and $`\delta T=TT_4`$ and vanish at $`(\mu _4,T_4)`$, while those of the quartic terms are
$`a_1^2={\displaystyle \frac{2B^2}{N_c}}{\displaystyle \frac{A^2N_c}{4}},b_1^2={\displaystyle \frac{A^2}{2}},c_1={\displaystyle \frac{2AB}{N_c}}{\displaystyle \frac{A^2}{2}},`$ (A5)
to leading order in $`\delta \mu `$ and $`\delta T`$.
The condition that the $`\chi \mathrm{\Delta }`$-phase includes $`(\mu _4,T_4)`$ can be determined as follows. While $`B/A`$ increases from below $`\alpha _2(N_c)`$, the tip of the wedge of the $`\chi \mathrm{\Delta }`$-phase slides along the the line of equal pressures between chiral and diquark phases, which we denote by $`L_{\chi \mathrm{\Delta },1}`$. Minimizing $`\mathrm{\Omega }(\sigma ,\mathrm{\Delta })`$ in Eq. (A4) to find the respective pressure in the $`\chi `$\- and $`\mathrm{\Delta }`$-phases, this line is determined near $`(\mu _4,T_4)`$ by
$`L_{\chi \mathrm{\Delta },1}{\displaystyle \frac{b_0^2}{b_1^2}}{\displaystyle \frac{a_0^2}{a_1^2}}=0.`$ (A6)
The left boundary of the $`\chi \mathrm{\Delta }`$-phase is the instability line
$`L_{\mathrm{mix}}{\displaystyle \frac{^2\mathrm{\Omega }}{\mathrm{\Delta }^2}}|_{\mathrm{\Delta }=0,\sigma =\sigma _0}=0,`$ (A7)
where $`\sigma _0`$ is the $`\sigma `$ field in the chiral phase. From Eq. (A4), this gives near $`(\mu _4,T_4)`$
$`L_{\mathrm{mix}}b_0{\displaystyle \frac{a_0}{a_1^2}}c_1=0.`$ (A8)
The tip $`P_{\mathrm{mix}}`$ of the $`\chi \mathrm{\Delta }`$-phase is the intercept of $`L_{\mathrm{mix}}`$ and $`L_{\chi \mathrm{\Delta },1}`$. From Eqs. (A6) and (A8) $`P_{\mathrm{mix}}`$ thus obeys
$`a_0^2\left(c_1^2a_1^2b_1^2\right)=0.`$ (A9)
Now, imagine that $`B/A`$ is strictly smaller but near $`\alpha _2(N_c)`$. Then, $`P_{\mathrm{mix}}(\mu _4,T_4)`$, and we have $`a_00`$. The location of $`P_{\mathrm{mix}}`$ is therefore determined by the condition $`a_1^2b_1^2=c_1^2`$, where the coefficients $`a_1`$, $`b_1`$ and $`c_1`$ are those of Eq. (A5) augmented by linear corrections in $`\delta \mu `$ and $`\delta T`$. If we now let $`B/A\alpha _2(N_c)`$, the tip $`P_{\mathrm{mix}}`$ reaches $`(\mu _4,T_4)`$ and the linear corrections in $`a_1`$, $`b_1`$, and $`c_1`$ vanish. Using Eq. (A5), and solving $`a_1^2b_1^2=c_1^2`$ for $`B/A`$ gives
$`\alpha _2(N_c)={\displaystyle \frac{B}{A}}{\displaystyle \frac{4N_cN_c^{3/2}\sqrt{2N_c4}}{4(4N_c)}},`$ (A10)
which is the result stated in Eq. (44).
Coming from large ratios $`B/A`$, the determining condition for $`B/A=\alpha _2(N_c)`$ requires that $`(\mu _4,T_4)`$ obeys Eq. (48),
$`{\displaystyle \frac{d^4\mathrm{\Omega }}{d\sigma ^4}}(\sigma ,\mathrm{\Delta }(\sigma ))`$ $`=`$ $`0,`$ (A11)
where $`\mathrm{\Delta }(\sigma )`$ is the diquark field in the $`\chi \mathrm{\Delta }`$-phase. Taking the derivative of Eq. (A4) with respect to $`\mathrm{\Delta }`$, we find $`\mathrm{\Delta }^2(\sigma )(c_1\sigma ^2+b_0)/b_1^2`$ which when inserted back in Eq. (A4) gives
$`{\displaystyle \frac{d^4\mathrm{\Omega }}{d\sigma ^4}}(\sigma ,\mathrm{\Delta }(\sigma ))`$ $``$ $`a_1^2{\displaystyle \frac{c_1^2}{b_1^2}},`$ (A12)
in the neighborhood of $`(\mu _4,T_4)`$. Setting the right side to zero, we recover the previous condition of Eq. (A9).
| Table 1. A few characteristic coupling ratios. | | | | | | | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- |
| Ratio $`B/A`$ | | $`N_c=2`$ | | | $`N_c=3`$ | | | $`N_c\mathrm{}`$ | | |
| Onset of diquark condensation | | | | | 0. | 418 | | 0. | 139 | $`N_c`$ |
| $`\alpha _1(N_c)`$ | | 1. | 0 | | 1. | 050 | | 0. | 321 | $`N_c`$ |
| Disappearance of the tricritical point. | | | | | 1. | 061 | | 0. | 354 | $`N_c`$ |
| $`\alpha _2(N_c)`$ | | 1. | 0 | | 1. | 163 | | 0. | 354 | $`N_c`$ |
| Maximum ratio $`N_c/2`$ | | 1. | 0 | | 1. | 5 | | 0. | 5 | $`N_c`$ |
| Single-gluon exchange | | 1. | 0 | | 0. | 75 | | 0. | 5 | | |
warning/0003/hep-th0003081.html | ar5iv | text | # Black hole pair creation in de Sitter space: a complete one-loop analysis
## 1 Introduction
Instantons play an important role in flat space gauge field theory . Being stationary points of the Euclidean action, they give the dominant contribution to the Euclidean path integral thus accounting for a variety of important phenomena in QCD-type theories. In addition, self-dual instantons admit supersymmetric extensions, which makes them an important tool for verifying various duality conjectures like the AdS/CFT correspondence . More generally, the Euclidean approach has become the standard method of quantum field theories in flat space.
Since the theory of gravity and Yang-Mills theory are somewhat similar, it is natural to study also gravitational instantons. An impressive amount of work has been done in this direction, leading to a number of important discoveries. A thorough study of instanton solutions of the vacuum Einstein equations and also those with a $`\mathrm{\Lambda }`$-term has been carried out . These solutions dominate the path integral of Euclidean quantum gravity, leading to interesting phenomena like black hole nucleation and quantum creation of universes. Perhaps one of the most spectacular achievements of the Euclidean approach is the derivation of black hole entropy from the action of the Schwarzschild instanton . In addition, gravitational instantons are used in the Kaluza-Klein reductions of string theory.
Along with these very suggestive results, the difficulties of Euclidean quantum gravity have been revealed. Apart from the usual problem of the non-renormalizability of gravity, which can probably be resolved only at the level of a more fundamental theory like string theory, the Euclidean approach presents other challenging problems. In field theories in flat space the correlation functions of field operators are holomorphic functions of the global coordinates in a domain that includes negative imaginary values of the time coordinate, $`t=i\tau `$, where $`\tau `$ is real and positive . This allows one to perform the analysis in the Euclidean section and then analytically continue the functions back to the Lorentzian sector to obtain the physical predictions. In curved space the theorems that would ensure the analyticity of any quantities arising in quantum gravity are not known. As a result, even if Euclidean calculations make sense, it is not in general clear how to relate their result to the Lorentzian physics.
This difficulty is most strikingly illustrated by the famous problem of the conformal sector in Euclidean quantum gravity. If one tries evaluating the path integral over Riemannian metrics, then one discovers that it diverges because the Euclidean gravitational action is not bounded from below and can be made arbitrarily large and negative by conformal rescaling of the metric . Such a result is actually expected, for if the integral did converge (with some regularization), then one could give a well-defined meaning to the canonical ensemble of the quantum gravitational field. However, the possibility of having a black hole causes the canonical ensemble to break down โ since the degeneracy of black hole states grows faster than the Boltzmann factor decreases. One can, โimproveโ the Euclidean gravitational action by analytically continuing the conformal modes, let us call them $`h`$, via $`hih`$, and this improves the convergence of the integral . This shows that if there is a well-defined Euclidean path integral for the gravitational field, then the relation to the Lorentzian sector is more complicated than just via $`ti\tau `$.
Unfortunately, it is unknown at present whether one can in the general case find a physically well-defined and convergent path integral for the gravitational field. At the same time, the idea of constructing it is conceptually simple : one should start from the Hamiltonian path integral over the physical degrees of freedom of the gravitational field. Such an integral certainly makes sense physically and is well-convergent, since the Hamiltonian is positive โ at least in the asymptotically flat case. The Hamiltonian approach is not covariant, but one can covariantize it by changing the integration variables, which leads to a manifestly covariant and convergent path integral for gravity. The main problem with this program is that in the general case it is unclear how to isolate the physical degrees of freedom of the gravitational field. For this reason, so far the program has been carried out only for weak fields in the asymptotically flat case . Remarkably, the result has been shown to exactly correspond to the the standard Euclidean path integral with the conformal modes being complex-rotated via $`hih`$. This lends support to the Euclidean approach in gravity and allows one to hope that the difficulties of the method can be consistently resolved; (see, for example, for the recent new developments within the lattice approach).
One can adopt the viewpoint that Euclidean quantum gravity is a meaningful theory within its range of applicability, at least at one-loop level, by assuming that a consistent resolution of its difficulties exists. Then already in its present status the theory can be used for calculating certain processes, most notably for describing tunneling phenomena, in which case the Euclidean amplitude directly determines the probability. The analytic continuation to the Lorentzian sector in this case is not necessary, apart from when the issue of the interpretation of the corresponding gravitational instanton is considered. The important example of a tunneling process in quantum gravity is the creation of black holes in external fields. Black holes are created whenever the energy pumped into the system is enough in order to make a pair of virtual black holes real . The energy can be provided by the heat bath , by the background magnetic field , by the expansion of the universe , by cosmic strings , domain walls , etc; (see also ). Besides, one can consider pair creation of extended multidimensional objects like $`p`$-branes due to interaction with the background supergravity fields . In all these examples the process is mediated by the corresponding gravitational instanton, and the semiclassical nucleation rate for a pair of objects on a given background is given by
$$\mathrm{\Gamma }=A\mathrm{exp}\left\{(I_{\mathrm{obj}}I_{\mathrm{bg}})\right\}.$$
(1.1)
Here $`I_{\mathrm{obj}}`$ is the classical action of the gravitational instanton mediating creation of the objects, $`I_{\mathrm{bg}}`$ is the action of the background fields alone, and the prefactor $`A`$ includes quantum corrections.
In most cases the existing calculations of black hole pair creation processes consider only the classical term in (1.1). This is easily understood, since loop calculations in quantum gravity for non-trivial backgrounds are extremely complicated. To our knowledge, there is only one example of a next-to-leading-order computation, which was undertaken in by Gross, Perry, and Yaffe for the Schwarzschild instanton background. The aim of the present paper is to consider one more example of a complete one-loop computation in quantum gravity.
The problem we are interested in is the quantum creation of black holes in de Sitter space. This problem was considered by Ginsparg and Perry , who identified the instanton responsible for this process, which is the $`S^2\times S^2`$ solution of the Euclidean Einstein equations $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$ for $`\mathrm{\Lambda }>0`$. Ginsparg and Perry noticed that this solution has one negative mode in the physical sector, which renders the partition function complex, thus indicating the quasi-classical instability of the system. This instability leads to spontaneous nucleation of black holes in the rapidly inflating universe. This is the dominant instability of de Sitter space, since classically the space is stable . The energy necessary for the nucleation is provided by the $`\mathrm{\Lambda }`$-term, which drives different parts of the universe apart thereby drugging the members of a virtual black hole pair away from each other. The typical radius of the created black holes is $`1/\sqrt{\mathrm{\Lambda }}`$, while the the nucleation rate is of the order of $`\mathrm{exp}(\pi /\mathrm{\Lambda }G)`$, where $`G`$ is Newtonโs constant. As a result, for $`\mathrm{\Lambda }G1`$ when inflation is fast, the black holes are produced in abundance but they are small and presumably almost immediately evaporate. Large black holes emerge for $`\mathrm{\Lambda }G1`$ when inflation slows down, and these can probably exist for a long time, but the probability of their creation is exponentially small. This scenario was further studied in Refs. (see also references in ), where the generalization to the charged case was considered and also the subsequent evolution of the created black holes was analyzed. However, the one-loop contribution so far has not been computed.
A remarkable feature of the $`S^2\times S^2`$ instanton is its high symmetry. In what follows, we shall utilize this symmetry in order to explicitly determine spectra of all relevant fluctuation operators in the problem. We shall use the $`\zeta `$-function regularization scheme in order to compute the one-loop determinants, which will give us the partition function $`Z[S^2\times S^2]`$ for the small fluctuations around the $`S^2\times S^2`$ instanton. We shall then need to normalize this result. The normalization coefficients is $`Z[S^4]`$, the partition function for small fluctuations around the $`S^4`$ instanton, which is the Euclidean version of the de Sitter space. The one-loop quantization around the $`S^4`$ instanton was considered by Gibbons and Perry , and by Christensen and Duff , but unfortunately in none of these papers the analysis was completed. We shall therefore reconsider the problem by rederiving the spectra of fluctuations around $`S^4`$ and computing the determinants within the $`\zeta `$-function scheme, thereby obtaining a closed one-loop expression for $`Z[S^4]`$.
In our treatment of the path integral we follow the approach of Gibbons and Perry ; (see also ). In order to have control over the results, we work in a one-parameter family of covariant gauges and perform the Hodge decomposition of the fluctuations. These are then expanded with respect to the complete sets of basis harmonics, and the perturbative path integration measure is defined as the square root of the determinant of the metric on the function space of fluctuations. To insure the convergence of the integral over the conformal modes, which enter the action with the wrong sign, we essentially follow the standard recipe $`hih`$ ; (see also Ref., where a slightly disguised form of the same prescription was advocated). The subtle issue is that the conformal operator $`\stackrel{~}{\mathrm{\Delta }}_0=3_\mu ^\mu 4\mathrm{\Lambda }`$ has a finite number, $`๐ฉ`$, of negative modes, and these enter the action with the correct sign from the very beginning. Our treatment of these special modes is different from that by Hawking , who suggests that such modes should be complex-rotated twice, the partition function then acquiring the overall factor of $`i^๐ฉ`$. However, the presence of this factor in the partition function would lead to unsatisfactory results, and on these grounds we are led to not rotating the special conformal modes at all.
The path integral is computed by integrating over the Fourier expansion coefficients, which leads to infinite products over the eigenvalues. The only conformal modes giving contribution to the result are the special negative modes discussed above. We carefully analyze the resulting products to make sure that all modes are taken into account and that the dependence of the gauge-fixing parameter cancels thereby indicating the correctness of the procedure. We give a detailed consideration to the zero modes of the Faddeev-Popov operator, which arise due to the background isometries. The integration over these modes requires a non-perturbative extension of the path-integration measure, and we find such a non-perturbative measure in the zero mode sector to be proportional to the Haar measure of the isometry group. Collecting all terms yields the partition function for small fluctuations around a background instanton configuration in terms of infinite products over eigenvalues of the gauge-invariant operators. We then use the explicitly known spectra of fluctuations around the $`S^2\times S^2`$ and $`S^4`$ backgrounds in order to calculate the partition functions.
The rest of the paper is organized as follows. In Sec.2 we present our derivation of the black hole nucleation rate within the finite temperature approach. In Sec.3 the path integration procedure is considered. The spectra of small fluctuations around the $`S^2\times S^2`$ instanton are computed in Sec.4 via a direct solving of the differential equations in the eigenvalue problems. The spectra of the fluctuations around the $`S^4`$ instanton are rederived in Sec.5 with the use of group theoretic arguments. The partition functions are computed in Sec.6, and Sec.7 contains the final expression for the black hole nucleation rate together with some remarks. We present a detailed analysis of the $`\zeta `$-functions in the Appendix. We use units where $`c=\mathrm{}=k_\mathrm{B}=1`$.
## 2 Black hole nucleation rate
In this section we shall derive the basic formula for the black hole nucleation rate in de Sitter space, whose different parts will be evaluated in the next sections. The existing derivations of the nucleation rate recover only the classical factor in (1.1). In addition, it is not always clear to which volume the rate refers. We argue that our formula (2.15) gives the nucleation probability per Hubble volume and unit time as measured by a freely falling observer. The basic idea of our approach is to utilize the relation between the inflation and thermal properties of de Sitter space. This will allow us to use the standard theory of decay of metastable thermal states .
Let us consider the partition function for the gravitational field
$$Z=D[g_{\mu \nu }]\mathrm{e}^I,$$
(2.2)
where the integral is taken over Riemannian metrics, and $`I=I[g_{\mu \nu }]`$ is the Euclidean action for gravity with a positive $`\mathrm{\Lambda }`$ terms; see Eq.(3.1) below. The path integration procedure will be considered in detail in the next section. At present let us only recall that in the semiclassical approximation the integral is approximated by the sum over the classical extrema of the action $`I`$, that is
$$Z\underset{l}{}Z_l.$$
(2.3)
Here $`Z_l=Z[_l]`$ is the partition function for the small gravitational fluctuations around a background manifold $`_l`$ with a metric $`g_{\mu \nu }^l`$ subject to the Euclidean Einstein equations $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$. Schematically one has
$$Z\underset{l}{}\frac{\mathrm{exp}(I_l)}{\sqrt{\mathrm{Det}\mathrm{\Delta }_l}},$$
(2.4)
where $`I_l`$ is the classical action for the $`l`$-th extremum, and $`\mathrm{\Delta }_l`$ is the operator for the small fluctuations around this background.
The dominant contribution to the sum in (2.4) is given by the $`S^4`$ instanton, which is the four-dimensional sphere with the radius $`\sqrt{3/\mathrm{\Lambda }}`$ and the standard metric. Since this is a maximally symmetry space, its action $`I=3\pi /\mathrm{\Lambda }G`$ is less than that of any other instanton. Hence,
$$ZZ[S^4]=\frac{\mathrm{exp}(3\pi /\mathrm{\Lambda }G)}{\sqrt{\mathrm{Det}\mathrm{\Delta }}}.$$
(2.5)
On the other hand, the $`S^4`$ instanton describes the thermal properties of de Sitter space , since it can be obtained by an analytic continuation via $`t\tau =it`$ of the region of the de Sitter solution
$$ds^2=(1\frac{\mathrm{\Lambda }}{3}r^2)dt^2+\frac{dr^2}{1\frac{\mathrm{\Lambda }}{3}r^2}+r^2(d\vartheta ^2+\mathrm{sin}^2\vartheta d\phi ^2)$$
(2.6)
contained inside the event horizon, $`r<\sqrt{3/\mathrm{\Lambda }}`$. Let us call this region a Hubble region. Its boundary, the horizon, has the area $`๐=12\pi /\mathrm{\Lambda }`$. The temperature associated with this horizon is $`T=\frac{1}{2\pi }\sqrt{\frac{\mathrm{\Lambda }}{3}}`$, the entropy $`S=๐/4G=3\pi /\mathrm{\Lambda }G`$ and the free energy $`F=TS`$. The same values can be obtained by writing the partition function for the $`S^4`$ instanton as
$$Z[S^4]=\mathrm{e}^{\beta F}$$
(2.7)
with $`\beta =1/T`$. Indeed, since $`S^4`$ is periodic in all four coordinates, any of them can be chosen to be the โimaginary timeโ. The corresponding period, $`\beta =2\pi \sqrt{\frac{3}{\mathrm{\Lambda }}}`$, can be identified with the proper length of a geodesic on $`S^4`$, all of which are circles with the same length. This gives the correct de Sitter temperature. Comparing (2.7) and (2.5) one obtains $`\beta F=3\pi /\mathrm{\Lambda }G+\mathrm{}`$ , the dots denoting the quantum corrections, and this again agrees with the result for the de Sitter space. To recapitulate, the partition function of quantum gravity with $`\mathrm{\Lambda }>0`$ is approximately
$$Z\mathrm{e}^{\beta F},$$
(2.8)
where $`1/\beta `$ is the de Sitter temperature and $`F`$ is the free energy in the Hubble region.
Let us now consider the contribution of the other instantons. One has
$$Z\mathrm{e}^{\beta F}\left(1+\underset{l}{\overset{}{}}\frac{Z[_l]}{Z[S^4]}\right),$$
(2.9)
where the prime indicates that $`_lS^4`$. Now, for $`\mathrm{\Lambda }G1`$ all terms in the sum are exponentially small and can safely be neglected as compared to the unity, if only they are real. If there are complex terms, then they will give an exponentially small imaginary contribution. The $`S^2\times S^2`$ instanton is distinguished by the fact that its partition function is purely imaginary due to the negative mode in the physical sector .
This is the only solution for $`\mathrm{\Lambda }>0`$ which is not a local minimum of the action in the class of metrics with constant scalar curvature . Hence (see Fig.1),
$$Z\mathrm{e}^{\beta F}\left(1+\frac{Z[S^2\times S^2]}{Z[S^4]}\right)\mathrm{exp}\left(\beta \left(F\frac{Z[S^2\times S^2]}{\beta Z[S^4]}\right)\right),$$
(2.10)
where $`Z[S^2\times S^2]`$ is purely imaginary. As a result, the partition function can still be represented as $`Z\mathrm{e}^{\beta F}`$, where the real part of $`F`$ is the free energy of the Hubble region, and the exponentially small imaginary part is given by
$$\mathrm{}(F)=\frac{Z[S^2\times S^2]}{\beta Z[S^4]}.$$
(2.11)
It is natural to relate this imaginary quantity also to the free energy. We are therefore led to the conclusion that the free energy of the Hubble region has a small imaginary part, thus indicating that the system is metastable. The decay of this metastable state will lead to a spontaneous nucleation of a black hole in the Hubble region, which can be inferred from the geometrical properties of the $`S^2\times S^2`$ instanton.
The $`S^2\times S^2`$ instanton can be obtained via the analytic continuation of the Schwarzschild-de Sitter solution
$$ds^2=Ndt^2+\frac{dr^2}{N}+r^2(d\vartheta ^2+\mathrm{sin}^2\vartheta d\phi ^2).$$
(2.12)
Here $`N=1\frac{2M}{r}\frac{\mathrm{\Lambda }}{3}r^2`$, and for $`9M^2\mathrm{\Lambda }<1`$ this function has roots at $`r=r_+>0`$ (black hole horizon) and at $`r=r_{++}>r_+`$ (cosmological horizon). One has $`N>0`$ for $`r_+<r<r_{++}`$, and only this portion of the solution can be analytically continued to the Euclidean sector via $`t\tau =it`$. The conical singularity at either of the horizons can be removed by a suitable identification of the imaginary time. However, since the two horizons have different surface gravities, the second conical singularity will survive. The situation improves in the extreme limit, $`r_+r_{++}\frac{1}{\sqrt{\mathrm{\Lambda }}}`$, since the surface gravities are then the same and both conical singularities can be removed at the same time. Although one might think that the Euclidean region shrinks to zero when the two horizons merge, this is not so. The limit $`r_+r_{++}`$ implies that $`9M^2\mathrm{\Lambda }=13ฯต^2`$ with $`ฯต0`$. One can introduce new coordinates $`\vartheta _1`$ and $`\phi _1`$ via $`\mathrm{cos}\vartheta _1=(\sqrt{\mathrm{\Lambda }}r1)/ฯต+ฯต/6`$ and $`\phi _1=\sqrt{\mathrm{\Lambda }}ฯต\tau `$. Passing to the new coordinates and taking the limit $`ฯต0`$, the result is
$$ds^2=\frac{1}{\mathrm{\Lambda }}(d\vartheta _{1}^{}{}_{}{}^{2}+\mathrm{sin}^2\vartheta _1d\phi _{1}^{}{}_{}{}^{2}+d\vartheta ^2+\mathrm{sin}^2\vartheta d\phi ^2),$$
(2.13)
and this $`S^2\times S^2`$ metric fulfills the Einstein equations. Since the instanton field determines the initial value for the created real time configuration, one concludes that the $`S^2\times S^2`$ instanton is responsible for the creation of a black hole in the Hubble region. This black hole fills the whole region, since its size is equal to the radius of the cosmological horizon.
It is well known that the region $`r<\sqrt{3/\mathrm{\Lambda }}`$ of the static coordinate system in (2.6) covers only a small portion of the de Sitter hyperboloid ; (see Fig.2). In order to cover the whole space, one can introduce an infinite number of freely falling observers and associate the interior of the static coordinate system with each of them. Hence, the spacetime contains infinitely many Hubble regions. It is also instructive to use global coordinates covering the whole de Sitter space,
$$ds^2=\frac{3}{\mathrm{\Lambda }\mathrm{cos}^2\xi }(d\xi ^2+d\chi ^2+\mathrm{sin}^2\chi (d\vartheta ^2+\mathrm{sin}^2\vartheta d\phi ^2)),$$
(2.14)
where $`\xi [\pi /2,\pi /2]`$ and $`\chi [0,\pi ]`$. The trajectory of a freely falling observer is $`\chi =\chi _0`$ (and also $`\vartheta =\vartheta _0`$, $`\phi =\phi _0`$), and the domain of the associated static coordinate system, the Hubble region, is the intersection of the interiors of the observerโs past and future horizons . Let $`\mathrm{\Sigma }`$ be a spacelike hypersurface, say $`\xi =\xi _0`$. If $`\xi _0=0`$ then $`\mathrm{\Sigma }`$ is completely contained inside the Hubble region of a single observer with $`\chi =\pi /2`$ (see Fig.2). However, for late moments of time, $`\xi \pi /2`$, one needs more and more independent observers in order to completely cover $`\mathrm{\Sigma }`$ by the union of their Hubble regions. One can say that the Hubble regions proliferate with the expansion of the universe.
Since de Sitter space consists of infinitely many Hubble regions, the black hole nucleation will lead to some of the regions being completely filled by a black hole, but most of the regions will be empty. The number of the filled regions divided by the number of those without a black hole is the probability for a black hole nucleation in one region. This is proportional to $`\mathrm{}(F)`$ in (2.11).
One can argue that the black holes are actually created in pairs , where the two members of the pair are located at the antipodal points of the de Sitter hyperboloid. This can be inferred from the conformal diagram of the Schwarzschild-de Sitter solution, which contains an infinite sequence of black hole singularities and spacelike infinities; see Fig.3. One can identify the asymptotically de Sitter regions in the diagram related by a horizontal shift, and the spacetime will then consist of two black holes at antipodal points of the closed universe. This agrees with the standard picture of particles in external fields being created in pairs.
The surface gravity of the extreme Schwarzschild-de Sitter solution is finite when defined with respect to the suitably normalized Killing vector . This gives a non-zero value for the temperature of the nucleated black holes, which can be read off also from the $`S^2\times S^2`$ metric: it is the inverse proper length of the equator of any of the two spheres, $`T_{\mathrm{BH}}=\frac{\sqrt{\mathrm{\Lambda }}}{2\pi }`$. How can it be that this is different from the temperature of the heat bath, which is the de Sitter space with $`T_{\mathrm{dS}}=\frac{1}{2\pi }\sqrt{\frac{\mathrm{\Lambda }}{3}}`$ ? For example, in the hot Minkowski space the nucleated black holes have the same temperature as the heat bath . However, the global structure of de Sitter space is different from that of Minkowski space. The fluctuations cannot absorb energy from and emit energy into the whole of de Sitter space, but can only exchange energy with the Hubble region. Thus the energy exchange is restricted. As a result, the local temperature in the vicinity of a created defect may be different from that of the heat bath, but reduces to the latter in the asymptotic region far beyond the cosmological horizon.
The relation of the imaginary part of the free energy to the rate of decay of a metastable thermal state $`\mathrm{\Gamma }`$ was considered in . If the decay is only due to tunneling then $`\mathrm{\Gamma }=2\mathrm{}(F)`$. Suppose that there is an additional possibility to classically jump over the potential barrier. In this case on top of the barrier there is a classical saddle point configuration whose real time decay rate is determined by the saddle negative mode $`\omega _{}`$. At low temperatures the tunneling formula is then still correct, while for $`T>\frac{|\omega _{}|}{2\pi }`$ one has $`\mathrm{\Gamma }=\frac{|\omega _{}|}{\pi T}\mathrm{}(F)`$. In our problem the saddle point configuration also exists, the $`S^2\times S^2`$ instanton, but its real time analog, the Schwarzschild-de Sitter black hole, is stable. It seems therefore that there is no classical contribution to the process and the black hole nucleation is a purely quantum phenomenon.<sup>1</sup><sup>1</sup>1We do not understand the classical interpretation of the Euclidean saddle point solution suggested in . The argument uses a family of non-normalizable deformations of the instanton, and the action is finite as long as they are โstaticโ. However, if one considers a time evolution along such a family then the action will be infinite, which shows that the classical picture does not apply. Even if one uses the classical formula for $`\mathrm{\Gamma }`$ in this case, one arrives at the quantum result, since $`|\omega _{}|/T`$=const.$`1`$. \[One can imagine that the effective potential barrier is infinitely high, such that a classical transition is forbidden, but at the same time so narrow that the tunneling rate is non-zero.\] As a result, the rate of quasiclassical decay of the de Sitter space is given by $`\mathrm{\Gamma }=2\mathrm{}(F)`$. Using Eq.(2.11),
$$\mathrm{\Gamma }=2T\frac{Z[S^2\times S^2]}{Z[S^4]}.$$
(2.15)
Here $`T=\frac{1}{2\pi }\sqrt{\frac{\mathrm{\Lambda }}{3}}`$ is the temperature of the de Sitter heat bath, which was originally defined with respect to the analytically continued Killing vector $`\frac{}{t}`$. Since $`t`$ is the proper time of the geodesic observer resting at the origin of the static coordinate system (2.6), we conclude that the formula gives the probability of a black hole nucleation per Hubble volume and unit time of a freely falling observer.
In order to use the formula (2.15), we should be able to compute the one-loop partition functions $`Z[S^2\times S^2]`$ and $`Z[S^4]`$. Now we shall calculate them within the path integral approach.
## 3 The path integration procedure
In this section we shall consider the path integral for fluctuations around an instanton solution of the Einstein equations $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$ in the stationary phase approximation. We shall largely follow the approach of Gibbons and Perry .
### 3.1 The second variation of the action
Our starting point is the action for the gravitational field on a compact Riemannian manifold $``$,
$$I[g_{\mu \nu }]=\frac{1}{16\pi G}_{}(R2\mathrm{\Lambda })\sqrt{g}d^4x,$$
(3.1)
whose extrema, $`\delta I=0`$, are determined by the equations
$$R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }.$$
(3.2)
Let $`g_{\mu \nu }`$ be an arbitrary solution, and consider small fluctuations around it, $`g_{\mu \nu }g_{\mu \nu }+h_{\mu \nu }`$. The action expands as
$$I[g_{\mu \nu }+h_{\mu \nu }]=I[g_{\mu \nu }]+\delta ^2I+\mathrm{},$$
(3.3)
where $`\delta ^2I`$ is quadratic in $`h_{\mu \nu }`$ and dots denote the higher order terms. One can express $`\delta ^2I`$ directly in terms if $`h_{\mu \nu }`$. However, it is convenient to use first the standard decomposition of $`h_{\mu \nu }`$,
$$h_{\mu \nu }=\varphi _{\mu \nu }+\frac{1}{4}hg_{\mu \nu }+_\mu \xi _\nu +_\nu \xi _\mu \frac{1}{2}g_{\mu \nu }_\sigma \xi ^\sigma .$$
(3.4)
Here $`\varphi _{\mu \nu }`$ is the transverse tracefree part, $`_\mu \varphi _\nu ^\mu =\varphi _\mu ^\mu =0`$, $`h`$ is the trace, and the piece due to $`\xi _\mu `$ is the longitudinal tracefree part. Under the gauge transformations (general diffeomorphisms) generated by $`\xi _\mu `$ one has $`h_{\mu \nu }h_{\mu \nu }+_\mu \xi _\nu +_\nu \xi _\mu .`$ The TT-tensor $`\varphi _{\mu \nu }`$ is gauge-invariant, while the trace $`h`$ changes as $`hh+2_\sigma \xi ^\sigma `$. It follows that
$$\stackrel{~}{h}=h2_\sigma \xi ^\sigma $$
(3.5)
is gauge-invariant. For further references we note that $`\xi _\mu `$ can in turn be decomposed into its coexact part $`\eta _\mu `$, for which $`_\mu \eta ^\mu =0`$, the exact part $`_\mu \chi `$, and the harmonic piece $`\xi _\mu ^\mathrm{H}`$,
$$\xi _\mu =\eta _\mu +_\mu \chi +\xi _\mu ^\mathrm{H}.$$
(3.6)
The number of square-integrable harmonic vectors is a topological invariant, which is equal to the first Betti number of the manifold $``$. Since the latter is zero if $``$ is simply-connected, which is the case for $`\mathrm{\Lambda }>0`$, we may safely ignore the harmonic contribution in what follows.
With the decomposition (3.4) the second variation of the action in (3.3) is expressed in terms of the gauge-invariant quantities $`\varphi _{\mu \nu }`$ and $`\stackrel{~}{h}`$ alone,
$$\delta ^2I=\frac{1}{2}\varphi ^{\mu \nu },\mathrm{\Delta }_2\varphi _{\mu \nu }\frac{1}{16}\stackrel{~}{h},\stackrel{~}{\mathrm{\Delta }}_0\stackrel{~}{h}.$$
(3.7)
Here and below we consider the following second order differential operators: the operator for the TT-tensor fluctuations
$$\mathrm{\Delta }_2\varphi _{\mu \nu }=_\sigma ^\sigma \varphi _{\mu \nu }2R_{\mu \alpha \nu \beta }\varphi ^{\alpha \beta },$$
(3.8)
the vector operator acting on coexact vectors $`\eta _\mu `$
$$\mathrm{\Delta }_1=_\sigma ^\sigma \mathrm{\Lambda },$$
(3.9)
and the scalar operators for $`h`$, $`\stackrel{~}{h}`$, and $`\chi `$
$`\mathrm{\Delta }_0`$ $`=`$ $`_\sigma ^\sigma ,`$
$`\stackrel{~}{\mathrm{\Delta }}_0`$ $`=`$ $`3\mathrm{\Delta }_04\mathrm{\Lambda },`$
$`\stackrel{~}{\mathrm{\Delta }}_0^\gamma `$ $`=`$ $`\gamma \stackrel{~}{\mathrm{\Delta }}_0\mathrm{\Delta }_0,`$ (3.10)
with $`\gamma `$ being a real parameter. Since for $`\mathrm{\Lambda }>0`$ the manifold $``$ is compact, these operators are (formally) self-adjoint with respect to the scalar product
$$\varphi _{\mu \nu },\varphi ^{\mu \nu }=\frac{1}{32\pi G}_{}\varphi _{\mu \nu }\varphi ^{\mu \nu }\sqrt{g}d^4x;$$
(3.11)
similarly for vectors $`\eta _\mu ,\eta ^\mu `$ and scalars $`\chi ,\chi `$.
The action $`\delta ^2I`$ in (3.7) contains only the gauge-invariant amplitudes $`\varphi _{\mu \nu }`$ and $`\stackrel{~}{h}`$, while the dependence on the gauge degrees of freedom $`\xi _\mu `$ cancels. Pure gauge modes are thus zero modes of the action. Fixing of the gauge is therefore necessary in order to carry out the path integration. To fix the gauge we pass from the action $`\delta ^2I`$ to the gauge-fixed action
$$\delta ^2I_{gf}=\delta ^2I+\delta ^2I_g,$$
(3.12)
where, following , we choose the gauge-fixing terms as
$$\delta ^2I_g=\gamma _\sigma h_\rho ^\sigma \frac{1}{\beta }_\rho h,^\alpha h_\alpha ^\rho \frac{1}{\beta }^\rho h,$$
(3.13)
with $`\gamma `$ and $`\beta `$ being real parameters. We shall shortly see that it is convenient to choose
$$\beta =\frac{4\gamma }{\gamma +1}.$$
(3.14)
This choice, however, implies that $`\delta ^2I_g`$ does not vanish for $`\gamma 0`$. It is often convenient to set $`\gamma =1`$, in which case $`\beta =2`$. However, we shall not fix the value of $`\gamma `$, since this will provide us with a check of the gauge-invariance of our results.
Using the decompositions (3.4), (3.6) the gauge-fixing term reads
$$\delta ^2I_g=\gamma \eta _\mu ,\mathrm{\Delta }_1^2\eta ^\mu +\frac{1}{16\gamma }(\stackrel{~}{h}+2\stackrel{~}{\mathrm{\Delta }}_0^\gamma \chi ),\mathrm{\Delta }_0(\stackrel{~}{h}+2\stackrel{~}{\mathrm{\Delta }}_0^\gamma \chi ).$$
(3.15)
Adding this up with $`\delta ^2I`$ in (3.7) one obtains the gauge-fixed action $`\delta ^2I_{gf}`$. It is now convenient to pass from the gauge-invariant variable $`\stackrel{~}{h}`$ defined in (3.5) back to the trace $`h`$, since with the choice in (3.14) the resulting action then becomes diagonal:
$`\delta ^2I_{gf}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\varphi ^{\mu \nu },\mathrm{\Delta }_2\varphi _{\mu \nu }+\gamma \eta _\mu ,\mathrm{\Delta }_1^2\eta ^\mu `$
$`+`$ $`{\displaystyle \frac{1}{4}}\chi ,\mathrm{\Delta }_0\stackrel{~}{\mathrm{\Delta }}_0\stackrel{~}{\mathrm{\Delta }}_0^\gamma \chi {\displaystyle \frac{1}{16\gamma }}h,\stackrel{~}{\mathrm{\Delta }}_0h.`$
This action generically has no zero modes, but it depends on the arbitrary parameter $`\gamma `$, which reflects the freedom of choice of gauge-fixing. In order to cancel this dependency, the compensating ghost term is needed.
### 3.2 The mode decomposition of the action
We wish to calculate the path integral
$$Z[g_{\mu \nu }]=\mathrm{e}^ID[h_{\mu \nu }]๐_{\mathrm{FP}}\mathrm{exp}\left(\delta ^2I_{gf}\right),$$
(3.17)
where $`I=I[g_{\mu \nu }]`$ is the classical action, and the Faddeev-Popov factor is obtained from
$$1=๐_{\mathrm{FP}}D[\xi _\mu ]\mathrm{exp}\left(\delta ^2I_g\right).$$
(3.18)
In order to perform the path integration, we introduce the eigenmodes associated with the operators $`\mathrm{\Delta }_2`$, $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_0`$:
$`\mathrm{\Delta }_2\varphi _{\mu \nu }^{(k)}`$ $`=`$ $`\epsilon _k\varphi _{\mu \nu }^{(k)},`$
$`\mathrm{\Delta }_1\eta _\mu ^{(s)}`$ $`=`$ $`\sigma _s\eta _\mu ^{(s)},`$
$`\mathrm{\Delta }_0\alpha ^{(p)}`$ $`=`$ $`\lambda _p\alpha ^{(p)}.`$ (3.19)
Throughout this paper we shall denote the eigenvalues and eigenfunctions of the tensor operator $`\mathrm{\Delta }_2`$ by $`\epsilon _k`$ and $`\varphi _{\mu \nu }^{(k)}`$, and those for the vector operator $`\mathrm{\Delta }_1`$ by $`\sigma _s`$ and $`\eta _\mu ^{(s)}`$, respectively. \[Later we shall use the symbol $`s`$ also for the argument of the $`\zeta `$-functions, and this will not lead to any confusion\]. Eigenvalues of the scalar operator will be denoted by $`\lambda _p`$, and it will be convenient to split the set $`\{\lambda _p\}`$ into three subsets, $`\{\lambda _p\}=\{\lambda _0,\lambda _i,\lambda _n\}`$, where $`\lambda _0=0`$, $`\lambda _i=\frac{4}{3}\mathrm{\Lambda }`$, and $`\lambda _n>\frac{4}{3}\mathrm{\Lambda }`$; see Eqs.(3.27)โ(3.27) below. Accordingly, the set of the scalar eigenfunctions will be split as $`\{\alpha ^{(p)}\}=\{\alpha ^{(0)},\alpha ^{(i)},\alpha ^{(n)}\}`$.
Since the manifold is compact, we choose the modes to be orthonormal. This allows us to expand all fields in the problem as
$$\varphi _{\mu \nu }=\underset{k}{}C_k^\varphi \varphi _{\mu \nu }^{(k)},\eta _\mu =\underset{s}{}C_s^\eta \eta _\mu ^{(s)},$$
(3.20)
and
$$\chi =\underset{p}{}C_p^\chi \alpha ^{(p)},h=\underset{p}{}C_p^h\alpha ^{(p)},\stackrel{~}{h}=\underset{p}{}C_p^{\stackrel{~}{h}}\alpha ^{(p)}.$$
(3.21)
As a result, the action decomposes into the sum over modes, and the path integral reduces to integrals over the Fourier coefficients.
a) Vector and tensor modes.โ Let us consider the mode decomposition for the gauge-fixed action in (3.1). This action is the sum of four terms. For the first two terms we obtain
$`{\displaystyle \frac{1}{2}}\varphi ^{\mu \nu },\mathrm{\Delta }_2\varphi _{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{k}{}}\epsilon _k(C_k^\varphi )^2,`$ (3.22)
$`\gamma \eta _\mu ,\mathrm{\Delta }_1^2\eta ^\mu `$ $`=`$ $`\gamma {\displaystyle \underset{s}{}}(\sigma _s)^2(C_s^\eta )^2.`$ (3.23)
These quadratic forms should be positive definite, since otherwise the integrals over the coefficients $`C`$ would be ill-defined. We can see that the quadratic form in (3.23) for the vector modes is indeed non-negative definite. Next, the expression in (3.22) for the gauge-invariant tensor modes is positive-definite if all eigenvalues $`\epsilon _k`$ are positive. If there is a negative eigenvalue, $`\epsilon _{}<0`$, as in the case of the $`S^2\times S^2`$ instanton background, then it is physically significant. The integration over $`C_{}^\varphi `$ is performed with the complex contour rotation, which renders the partition function imaginary thus indicating the quasiclassical instability of the system.
Let us consider now the contribution of the longitudinal vector piece to the action (3.1). We obtain
$$\frac{1}{4}\chi ,\mathrm{\Delta }_0\stackrel{~}{\mathrm{\Delta }}_0\stackrel{~}{\mathrm{\Delta }}_0^\gamma \chi =\frac{1}{4}\underset{p}{}\lambda _p\stackrel{~}{\lambda }_p\stackrel{~}{\lambda }_p^\gamma (C_p^\chi )^2,$$
(3.24)
where $`\stackrel{~}{\lambda }_p=3\lambda _p4\mathrm{\Lambda }`$ and $`\stackrel{~}{\lambda }_p^\gamma =\gamma \stackrel{~}{\lambda }_p\lambda _p`$ are the eigenvalues of $`\stackrel{~}{\mathrm{\Delta }}_0`$ and $`\stackrel{~}{\mathrm{\Delta }}_0^\gamma `$, respectively. We note that while $`\lambda _p0`$, the $`\stackrel{~}{\lambda }_p`$ and $`\stackrel{~}{\lambda }_p^\gamma `$ can be negative and should therefore be treated carefully. Let us split the scalar modes into three groups according to the sign of $`\stackrel{~}{\lambda }_p`$:
$`\mathrm{\Delta }_0\alpha ^{(0)}=0,\stackrel{~}{\lambda }_0={\displaystyle \frac{4}{3}}\mathrm{\Lambda },`$ (3.27)
$`\mathrm{\Delta }_0\alpha ^{(i)}={\displaystyle \frac{4}{3}}\mathrm{\Lambda }\alpha ^{(i)},\stackrel{~}{\lambda }_i=0,`$
$`\mathrm{\Delta }_0\alpha ^{(n)}=\lambda _n\alpha ^{(n)},\stackrel{~}{\lambda }_n>0.`$
First we consider the constant mode $`\alpha ^{(0)}`$ in (3.27). This exists for any background, and for compact manifolds without boundary this is the only normalizable zero mode of $`\mathrm{\Delta }_0`$. Since this mode is annihilated by $`\mathrm{\Delta }_0`$, it does not contribute to the sum in (3.24).
Consider now the scalar modes with the eigenvalue $`4\mathrm{\Lambda }/3`$ in (3.27). In view of the Lichnerowicz-Obata theorem , the lowest non-trivial eigenvalue of $`\mathrm{\Delta }_0`$ for $`\mathrm{\Lambda }>0`$ is bounded from below by $`4\mathrm{\Lambda }/3`$, and the equality is attained if only the background is $`S^4`$. Hence the modes in (3.27) exist only for the $`S^4`$ instanton, and there can be no modes โin betweenโ (3.27) and (3.27). In the $`S^4`$ case there are five scalar modes with the eigenvalue $`4\mathrm{\Lambda }/3`$, and their gradients are the five conformal Killing vectors that do not correspond to infinitesimal isometries. If $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$, a theorem of Yano an Nagano states that such vectors exist only in the $`S^4`$ case. Let us call these five scalar modes โconformal Killing modesโ. Notice that these also do not contribute to the sum in (3.24).
To recapitulate, the lowest lying modes in the scalar spectrum are the constant conformal mode in (3.27), which exists for any background, and also 5 โconformal Killing modesโ in (3.27) which exist only for the $`S^4`$ instanton and generate the conformal isometries. As we shall see, these 1+5 lowest lying modes are physically distinguished, since they are the only scalar modes contributing to the partition function. However, they do not enter the sum in (3.24).
For the remaining infinite number of scalar modes in (3.27) (these are labeled by $`n`$) the eigenvalues $`\lambda _n`$ and $`\stackrel{~}{\lambda }_n`$ are positive, and it is not difficult to see that all the $`\stackrel{~}{\lambda }_n^\gamma `$โs are also positive, provided that the gauge parameter $`\gamma `$ is positive and large enough. To recapitulate, the contribution of the longitudinal vector modes to the action is given by
$$\frac{1}{4}\chi ,\mathrm{\Delta }_0\stackrel{~}{\mathrm{\Delta }}_0\stackrel{~}{\mathrm{\Delta }}_0^\gamma \chi =\frac{1}{4}\underset{n}{}\lambda _n\stackrel{~}{\lambda }_n\stackrel{~}{\lambda }_n^\gamma (C_n^\chi )^2,$$
(3.28)
which is positive definite. We shall see that all modes contributing to this sum are unphysical and cancel from the path integral.
b) Conformal modes.โ We now turn to the last term in the gauge-fixed action (3.1). Using (3.27)โ(3.27) we obtain
$$\frac{1}{16\gamma }h,\stackrel{~}{\mathrm{\Delta }}_0h=\frac{\mathrm{\Lambda }}{4}(C_0^h)^2+\frac{\mathrm{\Lambda }}{12\gamma }\underset{i}{}(C_i^h)^2\frac{1}{16\gamma }\underset{n}{}\stackrel{~}{\lambda }_n^\gamma (C_n^h)^2.$$
(3.29)
The expression on the right has a finite number of positive terms, corresponding to the distinguished lowest lying modes, and infinitely many negative ones. As a result, an increase in the coefficients $`C_n^h`$ makes it arbitrarily large and negative, thus rendering the path integral divergent. This represents the well-known problem of conformal modes in Euclidean quantum gravity . A complete solution of this problem is lacking at present, but the origin of the trouble seems to be understood . In brief, the problem is not related to any defects of the theory itself, but arises as a result of the bad choice of the path integral. If one starts from the fundamental Hamiltonian path integral over the physical degrees of freedom of the gravitational field, then one does not encounter this problem. The Hamiltonian path integral, however, is non-covariant and difficult to work with. One can โcovariantizeโ it by adding gauge degrees of freedom, and this leads to the Euclidean path integral described above, up to the important replacement
$$hih.$$
(3.30)
The effect of this is to change the overall sign in (3.29), such that the infinite number of negative modes become positive. Unfortunately, such a consistent derivation of the path integral has only been carried out for weak gravitational fields (and for $`\mathrm{\Lambda }=0`$), since otherwise it is unclear how to choose the physical degrees of freedom. Nevertheless, the rule (3.30) is often used also in the general case , and it leads to the cancellation of the unphysical conformal modes. However, some subtle issues can arise.
For $`\mathrm{\Lambda }>0`$ the expression in (3.29) contains, apart from infinitely many negative terms, also a finite number of positive ones, which are due to the distinguished lowest lying scalar modes. If we apply the rule (3.30) and change the overall sign of the scalar mode action, then the negative modes will become positive, but the positive ones will become negative. As a result, the path integral will still be divergent. It was therefore suggested by Hawking that the contour for these extra negative modes should be rotated back, the partition function then acquiring the factor $`i^๐ฉ`$, where $`๐ฉ`$ is the number of such modes . As we know, $`๐ฉ=6`$ for the $`S^4`$ instanton, and $`๐ฉ=1`$ for any other solution of $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$ with $`\mathrm{\Lambda }>0`$.
Unfortunately, this prescription to rotate the contour twice leads in some cases to physically meaningless results; the examples will be given in a moment. We suggest therefore a slightly different scheme: not to touch the positive modes in (3.29) at all and to change the sign only for the negative modes. The whole expression then becomes
$$\frac{1}{16\gamma }h,\stackrel{~}{\mathrm{\Delta }}_0h=\frac{\mathrm{\Lambda }}{4}(C_0^h)^2+\frac{\mathrm{\Lambda }}{12\gamma }\underset{i}{}(C_i^h)^2+\frac{1}{16\gamma }\underset{n}{}\stackrel{~}{\lambda }_n^\gamma (C_n^h)^2.$$
(3.31)
We make no attempt to rigorously justify such a rule. We note, however, that it is essentially equivalent to the standard recipe (3.30) โ up to a finite number of modes which we handle differently as compared to Hawkingโs prescription. We shall now comment on this difference.
When compared to Hawkingโs recipe , the ultimate effect of our prescription is to remove the factor $`i^๐ฉ`$ from the partition function. We are unaware of any examples where it would be necessary to insist on this factor being present in the final result. On the contrary, the examples are in favour of the factor being absent. For the $`S^4`$ instanton one has $`๐ฉ=6`$, such that $`i^๐ฉ=1`$, and this would render the partition function for hot gravitons in a de Sitter universe negative, which would be physically meaningless. Next, for the $`S^2\times S^2`$ instanton, which already has one negative mode in the spin-2 sector, one has $`๐ฉ=1`$. As a result, the factor $`i^๐ฉ`$ would make the partition function real instead of being imaginary, and there would be no black hole pair creation !
These arguments suggest that Hawkingโs rule should be somehow modified, and we therefore put forward the prescription (3.31). Let us also note that our rule leads to gauge invariant results โ the dependence on the gauge parameter $`\gamma `$ cancels after the integration. Finally we note that the lowest lying scalar modes are physically distinguished, and since they are positive, they should be treated similarly to the physical tensor modes.
To recapitulate, the mode expansion of the gauge-fixed action $`\delta ^2I_{gf}`$ is given by the sum of (3.22), (3.23), (3.28), and (3.31):
$`\delta ^2I_{gf}={\displaystyle \frac{1}{2}}{\displaystyle \underset{k}{}}\epsilon _k(C_k^\varphi )^2+\gamma {\displaystyle \underset{s}{}}(\sigma _s)^2(C_s^\eta )^2+{\displaystyle \frac{1}{4}}{\displaystyle \underset{n}{}}\lambda _n\stackrel{~}{\lambda }_n\stackrel{~}{\lambda }_n^\gamma (C_n^\chi )^2`$
$`+{\displaystyle \frac{\mathrm{\Lambda }}{4}}(C_0^h)^2+{\displaystyle \frac{\mathrm{\Lambda }}{12\gamma }}{\displaystyle \underset{i}{}}(C_i^h)^2+{\displaystyle \frac{1}{16\gamma }}{\displaystyle \underset{n}{}}\stackrel{~}{\lambda }_n^\gamma (C_n^h)^2.`$ (3.32)
In a similar way we obtain the following mode expansion for the gauge-fixing term $`\delta ^2I_g`$ in (3.15):
$`\delta ^2I_g`$ $`=`$ $`\gamma {\displaystyle \underset{s}{}}(\sigma _s)^2(C_s^\eta )^2+{\displaystyle \frac{16}{27\gamma }}\mathrm{\Lambda }^3{\displaystyle \underset{i}{}}\left(C_i^\chi {\displaystyle \frac{3}{8\mathrm{\Lambda }}}C_i^{\stackrel{~}{h}}\right)^2`$ (3.33)
$`+`$ $`{\displaystyle \frac{1}{16\gamma }}{\displaystyle \underset{n}{}}\lambda _n(2\stackrel{~}{\lambda }_n^\gamma C_n^\chi +C_n^{\stackrel{~}{h}})^2.`$
This expression is non-negative definite.
### 3.3 The path integration measure
In order to compute the path integrals in (3.17),(3.18) we still need to define the path-integration measure. The perturbative measure is defined as the square root of the determinant of the metric on the function space of fluctuations:
$$D[h_{\mu \nu }]\sqrt{\mathrm{Det}(dh_{\mu \nu },dh^{\mu \nu }}),D[\xi _\mu ]\sqrt{\mathrm{Det}(d\xi _\mu ,d\xi ^\mu }).$$
(3.34)
Here it is assumed that the fluctuations are Fourier-expanded and the differentials refer to the Fourier coefficients, while the meaning of the proportionality sign will become clear shortly. Let us first consider $`D[\xi _\mu ]`$. It follows from (3.4),(3.6) that
$`h_{\mu \nu },h^{\mu \nu }`$ $`=`$ $`\varphi _{\mu \nu },\varphi ^{\mu \nu }+2\eta _\mu ,\mathrm{\Delta }_1\eta ^\mu +\chi ,\mathrm{\Delta }_0\stackrel{~}{\mathrm{\Delta }}_0\chi +{\displaystyle \frac{1}{4}}h,h,`$
$`\xi _\mu ,\xi ^\mu `$ $`=`$ $`\eta _\mu ,\eta ^\mu +\chi ,\mathrm{\Delta }_0\chi .`$ (3.35)
Expanding the fields on the right according to (3.20),(3.21) and differentiating with respect to the Fourier coefficients we obtain the metric for the vector fluctuations
$`d\xi _\mu ,d\xi ^\mu =d\eta _\mu ,d\eta ^\mu +d\chi ,\mathrm{\Delta }_0d\chi ={\displaystyle \underset{s}{}}(dC_s^\eta )^2+{\displaystyle \underset{p}{\overset{}{}}}\lambda _p(dC_p^\chi )^2,`$ (3.36)
which yields
$$\sqrt{\mathrm{Det}(d\xi _\mu ,d\xi ^\mu )}=\left(\underset{s}{}dC_s^\eta \right)\left(\underset{p}{\overset{}{}}\sqrt{\lambda _p}dC_p^\chi \right).$$
(3.37)
Here the prime indicates that terms with $`\lambda _p=0`$ do not contribute to the sum in (3.36), and should therefore be omitted in the product in (3.37). To obtain the measure $`D[\xi _\mu ]`$ we endow each term in the products in (3.37) with the weight factor $`\mu _o^2/\sqrt{\pi }`$:
$$D[\xi _\mu ]=\left(\underset{s}{}\frac{\mu _o^2}{\sqrt{\pi }}dC_s^\eta \right)\left(\underset{i}{}\frac{\mu _o^2}{\sqrt{\pi }}\sqrt{\frac{4\mathrm{\Lambda }}{3}}dC_i^\chi \right)\left(\underset{n}{}\frac{\mu _o^2}{\sqrt{\pi }}\sqrt{\lambda _n}dC_n^\chi \right).$$
(3.38)
Such a normalization implies that
$$1=D[\xi _\mu ]\mathrm{exp}\left(\mu _o^4\xi _\mu ,\xi ^\mu \right).$$
(3.39)
Here $`\mu _o`$ is a parameter with the dimension of an inverse length. In a similar way we obtain the measure $`D[h_{\mu \nu }]`$, which is normalized as
$$1=D[h_{\mu \nu }]\mathrm{exp}\left(\frac{\mu _o^2}{2}h_{\mu \nu },h^{\mu \nu }\right);$$
(3.40)
we shall shortly comment on the relative normalization of $`D[h_{\mu \nu }]`$ and $`D[\xi _\mu ]`$. The result is
$`D[h_{\mu \nu }]`$ $`=`$ $`\left({\displaystyle \underset{k}{}}{\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}dC_k^\varphi \right)\left({\displaystyle \underset{s}{\overset{}{}}}{\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}\sqrt{2\sigma _s}dC_s^\eta \right)\left({\displaystyle \underset{n}{}}{\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}\sqrt{\lambda _n\stackrel{~}{\lambda }_n}dC_n^\chi \right)`$ (3.41)
$`\times `$ $`\left({\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}{\displaystyle \frac{1}{2}}dC_0^h\right)\left({\displaystyle \underset{i}{}}{\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}{\displaystyle \frac{1}{2}}dC_i^h\right)\left({\displaystyle \underset{n}{}}{\displaystyle \frac{\mu _o}{\sqrt{2\pi }}}{\displaystyle \frac{1}{2}}dC_n^h\right).`$
Here the prime indicates that the zero modes of the vector fluctuation operator do not contribute to the product. Notice, however, that these modes do contribute to the measure $`D[\xi _\mu ]`$.
The following remarks are in order. We use units where all fields and parameters have dimensions of different powers of a length scale $`l`$. One has $`[1/G]=[\mathrm{\Lambda }]=[\mu _o^2]=[l^2]`$. Eigenvalues of all fluctuation operators have the dimension $`[l^2]`$. The coordinates $`x^\mu `$ are dimensionless, while $`[g_{\mu \nu }]=[h_{\mu \nu }]=[l^2]`$. For the vectors, $`[\eta _\mu ]=[\xi _\mu ]=[l^2]`$, and for the scalars $`[h]=[l^0]`$ and $`[\chi ]=[l^2].`$ We assume that the scalar, vector and tensor eigenfunctions in (3.19) are orthonormal with respect to the scalar product in (3.11). As a result, the dimensions of the eigenfunctions are $`[\varphi _{\mu \nu }^{(k)}]=[l]`$, $`[\eta _\mu ^{(s)}]=[l^0]`$, $`[\alpha ^{(p)}]=[l^1]`$, which gives for the Fourier coefficients in (3.20),(3.21) $`[C^\varphi ]=[C^h]=[l]`$, $`[C^\eta ]=[l^2]`$, and $`[C^\chi ]=[l^3]`$.
The normalization of $`D[h_{\mu \nu }]`$ can be arbitrary, which is reflected in the presence of the arbitrary parameter $`\mu _o`$ in the above formulas. However, the relative normalization of $`D[h_{\mu \nu }]`$ and $`D[\xi _\mu ]`$, which is defined by Eqs.(3.39) and (3.40) is fixed by gauge invariance. Had we chosen instead a different relative normalization, say dividing each mode in (3.38) by 2, then the path integral would acquire a factor of $`2^{\stackrel{~}{N}_0^\gamma }`$, where $`\stackrel{~}{N}_0^\gamma `$ is the โnumber of eigenvaluesโ of the non-gauge-invariant operator $`\stackrel{~}{\mathrm{\Delta }}_0^\gamma `$. \[The issue of relative normalization of the fluctuation and Faddeev-Popov determinants seldom arises, since in most cases the absolute value of the path integral is irrelevant\].
### 3.4 Computation of the path integral
Now we are ready to compute the path integrals in (3.17),(3.18). Let us illustrate the procedure on the example of Eq.(3.18), which determines the Faddeev-Popov factor $`๐_{FP}`$. Using $`\delta ^2I_g`$ from Eq.(3.33) and the measure $`D[\xi _\mu ]`$ from (3.38) we obtain
$`(๐_{FP})^1={\displaystyle \underset{s}{}}{\displaystyle \frac{\mu _o^2}{\sqrt{\pi }}๐C_s^\eta \mathrm{exp}\left(\gamma (\sigma _s)^2(C_s^\eta )^2\right)}`$ (3.42)
$`\times {\displaystyle \underset{i}{}}{\displaystyle }{\displaystyle \frac{\mu _o^2}{\sqrt{\pi }}}\sqrt{{\displaystyle \frac{4\mathrm{\Lambda }}{3}}}dC_i^\chi \mathrm{exp}({\displaystyle \frac{16}{27\gamma }}\mathrm{\Lambda }^3(C_i^\chi {\displaystyle \frac{3}{8\mathrm{\Lambda }}}C_i^{\stackrel{~}{h}})^2)`$
$`\times {\displaystyle \underset{n}{}}{\displaystyle }{\displaystyle \frac{\mu _o^2}{\sqrt{\pi }}}\sqrt{\lambda _n}dC_n^\chi \mathrm{exp}({\displaystyle \frac{1}{16\gamma }}{\displaystyle \underset{n}{}}\lambda _n(2\stackrel{~}{\lambda }_n^\gamma C_n^\chi +C_n^{\stackrel{~}{h}})^2),`$
which gives
$$(๐_{FP})^1=\mathrm{\Omega }_1\left(\underset{s}{\overset{}{}}\frac{\mu _o^2}{\sqrt{\gamma }\sigma _s}\right)\left(\underset{i}{}\frac{3\sqrt{\gamma }\mu _o^2}{2\mathrm{\Lambda }}\right)\left(\underset{n}{}\frac{2\sqrt{\gamma }\mu _o^2}{\stackrel{~}{\lambda }_n^\gamma }\right).$$
(3.43)
#### 3.4.1 Zero modes
The factor $`\mathrm{\Omega }_1`$ in (3.43) arises due to the gauge zero modes, for which $`\sigma _s\sigma _0=0`$ and the integral is non-Gaussian:
$$\mathrm{\Omega }_1=\underset{j}{}\frac{\mu _o^2}{\sqrt{\pi }}dC_{0j}^\eta ,$$
(3.44)
with the product taken over all such modes. The existence of zero modes of the Faddeev-Popov operator indicates that the gauge is not completely fixed. This can be related to the global aspects of gauge fixing procedure known as the Gribov ambiguity. However, Gribovโs problem is usually not the issue in the perturbative calculations, where zero modes arise rather due to background symmetries. This will be the case in our analysis below. Specifically, the isometries of the background manifold $``$ form a subgroup $``$ of the full diffeomorphism group. Sometimes $``$ is called the stability group; for the $`S^4`$ and $`S^2\times S^2`$ backgrounds $``$ is SO(5) and SO(3)$`\times `$SO(3), respectively. Since the isometries do not change $`h_{\mu \nu }`$ (in the linearized approximation), their generators, which are the Killing vectors $`K^\mu `$, are zero modes of the Faddeev-Popov operator.
We therefore conclude that the integration in (3.44) is actually performed over the stability group $``$. Since the latter is compact in the cases under consideration, the integral is finite. In order to actually compute the integral, some further analysis is necessary, in which we shall adopt the approach of Osborn . First of all, let us recall that all eigenmodes in our analysis have unit norm. If we now rescale the zero modes such that the Killing vectors $`K_jK_j^\mu \frac{}{x^\mu }`$ become dimensionless (remember that the coordinates $`x^\mu `$ are also dimensionless), then the expression in Eq.(3.44) reads
$$\mathrm{\Omega }_1=\underset{j}{}\frac{\mu _o^2}{\sqrt{\pi }}K_jdC_j,$$
(3.45)
where now $`[K_j]=[l^2]`$ and $`[C_j]=[l^0]`$. For small values of the parameters $`C_j`$ they can be regarded as coordinates on the group manifold $``$ in the vicinity of the unit element. Since $``$ acts on $``$ via $`x^\mu x^\mu (C_j)`$, one has $`K_j=\frac{}{C_j}\frac{x^\mu }{C_j}\frac{}{x^\mu }`$. However, strictly speaking $`C_j`$ are not coordinates on the group manifold $``$ but rather on its tangent space at the group unity, such that their range is infinite. We wish to restrict the range of $`C_j`$, and for this we should integrate not over the tangent space but over $``$ itself. In other words, to render the integral in (3.45) convergent we must treat the zero modes non-perturbatively, and for this we should replace the perturbative measure $`_jdC_j`$ by a non-perturbative one, $`d\mu (C)`$.
In general it is a difficult issue to construct the non-perturbative path integration measure. However, in the zero mode sector this can be done. We note that the measure should be invariant under the group multiplications, $`d\mu (CC^{})=d\mu (C)`$, and this uniquely requires that $`d\mu (C)`$ should be the Haar measure for $``$. The normalization is fixed by the requirement that for $`C_j0`$ the perturbative result (3.45) is reproduced. This unambiguously gives
$$\mathrm{\Omega }_1=\left(\underset{j}{}\frac{\mu _o^2}{\sqrt{\pi }}\frac{}{C_j}\right)๐\mu (C),$$
(3.46)
where $`\frac{}{C_j}`$ is computed at $`C_j=0`$ and $`d\mu (C)`$ is the Haar measure of the isometry group $``$ normalized such that $`d\mu (C)_jdC_j`$ as $`C_j0`$.
#### 3.4.2 The path integral
Now, using (3.2) and the measure (3.41), we compute the path integral in (3.17) โ first without the Faddeev-Popov factor $`๐_{FP}`$:
$`{\displaystyle D[h_{\mu \nu }]\mathrm{exp}\left(\delta ^2I_{gf}\right)}`$ $`=`$ $`\mathrm{\Omega }_2\left({\displaystyle \underset{k}{\overset{}{}}}{\displaystyle \frac{\mu _o}{\sqrt{ฯต_k}}}\right)\left({\displaystyle \underset{s}{\overset{}{}}}{\displaystyle \frac{\mu _o}{\sqrt{\gamma \sigma _s}}}\right)\left({\displaystyle \underset{n}{}}{\displaystyle \frac{\sqrt{2}\mu _o}{\sqrt{\stackrel{~}{\lambda }_n^\gamma }}}\right)`$ (3.47)
$`\times `$ $`{\displaystyle \frac{\mu _o}{\sqrt{2\mathrm{\Lambda }}}}\left({\displaystyle \underset{i}{}}{\displaystyle \frac{\sqrt{3\gamma }\mu _o}{\sqrt{2\mathrm{\Lambda }}}}\right)\left({\displaystyle \underset{n}{}}{\displaystyle \frac{\sqrt{2\gamma }\mu _o}{\sqrt{\stackrel{~}{\lambda }_n^\gamma }}}\right).`$
Here the primes indicate that zero and negative modes should be omitted from the products. Zero vector modes do not contribute since they are not present in the path-integration measure (3.41), and we assume that there are no negative vector modes, since otherwise the metric on the space of fluctuations would not be positive definite. For tensor fluctuations negative and zero modes are present in the measure (3.41), and their overall contribution is collected in the factor $`\mathrm{\Omega }_2`$ in (3.47). Let us further assume that there are no zero tensor modes, which is the case for the manifolds of interest. If negative modes are also absent then $`\mathrm{\Omega }_2=1`$. If there is one negative tensor mode with eigenvalue $`\epsilon _{}<0`$, then
$$\mathrm{\Omega }_2=\frac{\mu _o}{\sqrt{2\pi }}๐C_{}^\varphi \mathrm{exp}\left(\frac{1}{2}\epsilon _{}(C_{}^\varphi )^2\right).$$
(3.48)
The integral is computed via the deformation of the contour to the complex plane, which gives the purely imaginary result
$$\mathrm{\Omega }_2=\frac{\mu _o}{2i\sqrt{|\epsilon _{}|}},$$
(3.49)
with the factor of 1/2 arising in the course of the analytic continuation .
Both the Faddeev-Popov factor in (3.43) and the path integral in (3.47) depend on the gauge parameter $`\gamma `$. However, the $`\gamma `$-dependence exactly cancels in their product, which provides a very good consistency check. In particular, the relative normalization of the integration measures fixed by Eqs.(3.39) and (3.40) is important. If we had divided each factor in the mode products in (3.38) by $`a1`$, then the resulting path integral would be proportional to $`(_na)a^{\zeta (0)}`$ with $`\zeta `$ being the $`\zeta `$-function of the $`\gamma `$-dependent operator $`\stackrel{~}{\mathrm{\Delta }}_0^\gamma `$. Thus, unless $`a=1`$, the result would be gauge-dependent.
We therefore finally obtain the following expression for the path integral in (3.17):
$$Z[g_{\mu \nu }]=\frac{\mathrm{\Omega }_2}{\mathrm{\Omega }_1}\frac{\mu _o}{\sqrt{2\mathrm{\Lambda }}}\left(\underset{i}{}\sqrt{\frac{2\mathrm{\Lambda }}{3}}\frac{1}{\mu _o}\right)\left(\underset{s}{\overset{}{}}\frac{\sqrt{\sigma _s}}{\mu _o}\right)\left(\underset{k}{\overset{}{}}\frac{\mu _o}{\sqrt{ฯต_k}}\right)\mathrm{e}^I$$
(3.50)
Here $`\mathrm{\Omega }_2`$ is the contribution of the negative tensor mode, and $`\mathrm{\Omega }_1`$ is the isometry factor. As we expected, the contribution of all unphysical scalar modes has canceled from the result. The only scalar modes which do contribute are the several lowest lying modes which seem to be physically distinguished. These are the constant conformal mode giving rise to the factor $`\mu _o/\sqrt{2\mathrm{\Lambda }}`$, and the 5 โconformal Killing scalarsโ which exist only in the $`S^4`$ case and give rise to the product over $`i`$. The next two factors in (3.50) is the contribution of the transversal vector modes and the TT-tensor modes. Finally, $`I=I[g_{\mu \nu }]`$ is the classical action.
In order to apply the above formula for $`Z[g_{\mu \nu }]`$ we need the eigenvalues of the fluctuation operators. Now we shall determine the latter for the manifolds $`S^2\times S^2`$ and $`S^4`$.
## 4 Spectra of fluctuation operators
In this section we derive the spectra of small fluctuations around the $`S^2\times S^2`$ and $`S^4`$ instantons. In the $`S^2\times S^2`$ case the problem is tackled via solving the differential equations. It turns out that in a suitable basis the system of 10 coupled equations for the gravity fluctuations splits into 10 independent equations. The latter are solved in terms of spin-weighted spherical harmonics. In the $`S^4`$ case the equations do not decouple and the direct approach is not so transparent. However, the problem can be conveniently analyzed with group theoretic methods, which was done some time ago by Gibbons and Perry . We shall describe the group theory approach in some detail โ with the same principal result as in .
### 4.1 Fluctuations around the $`S^2\times S^2`$ instanton
Let us consider the metric of the $`S^2\times S^2`$ instanton background,
$$ds^2=\frac{1}{\mathrm{\Lambda }}((d\vartheta _1)^2+\mathrm{sin}^2\vartheta _1(d\phi _1)^2+(d\vartheta _2)^2+\mathrm{sin}^2\vartheta _2(d\phi _2)^2).$$
(4.1)
It is convenient to set $`\mathrm{\Lambda }=1`$ for the time being; at the end of calculations the $`\mathrm{\Lambda }`$-dependence is restored by multiplying all eigenvalues with $`\mathrm{\Lambda }`$. We introduce the complex tetrad
$`e^1={\displaystyle \frac{1}{\sqrt{2}}}(d\vartheta _1+{\displaystyle \frac{i}{\mathrm{sin}\vartheta _1}}d\phi _1),e^2={\displaystyle \frac{1}{\sqrt{2}}}(d\vartheta _1{\displaystyle \frac{i}{\mathrm{sin}\vartheta _1}}d\phi _1),`$
$`e^3={\displaystyle \frac{1}{\sqrt{2}}}(d\vartheta _2+{\displaystyle \frac{i}{\mathrm{sin}\vartheta _2}}d\phi _2),e^4={\displaystyle \frac{1}{\sqrt{2}}}(d\vartheta _2{\displaystyle \frac{i}{\mathrm{sin}\vartheta _2}}d\phi _2).`$ (4.2)
The metric in (4.1) splits as $`g_{\mu \nu }=e_\mu ^ae_\nu ^b\eta _{ab}`$, where the tetrad metric is
$$\eta ^{ab}=g^{\mu \nu }e_\mu ^ae_\nu ^b=\left(\begin{array}{cccc}0& 1& 0& 0\\ 1& 0& 0& 0\\ 0& 0& 0& 1\\ 0& 0& 1& 0\end{array}\right).$$
(4.3)
#### 4.1.1 Tensor modes
First we consider the eigenvalue problem
$$_\alpha ^\alpha \varphi _{\mu \nu }2R_{\mu \alpha \nu \beta }\varphi ^{\alpha \beta }=\epsilon \varphi _{\mu \nu },$$
(4.4)
where
$$_\mu \varphi _\nu ^\mu =0,\varphi _\mu ^\mu =0.$$
(4.5)
We expand $`\varphi _{\mu \nu }`$ with respect to the complex basis (4.1),
$$\varphi _{\mu \nu }=e_\mu ^ae_\nu ^b\mathrm{\Phi }_{ab},$$
(4.6)
insert this into (4.4) and project the resulting equation onto the basis (4.1) again. Remarkably, the system of 10 coupled equations splits then into 10 independent equations for the 10 tetrad projections $`\mathrm{\Phi }_{ab}`$. A partial explanation of this fact is the existence of two different parity symmetries acting independently on the two spheres.
It is convenient to introduce the operator
$$\widehat{๐}[๐ฌ,\vartheta ,\phi ]=\frac{^2}{\vartheta ^2}+\mathrm{cot}\vartheta \frac{}{\vartheta }+2i๐ฌ\frac{\mathrm{cot}\vartheta }{\mathrm{sin}\vartheta }+\frac{1}{\mathrm{sin}^2\vartheta }\frac{^2}{\varphi ^2}๐ฌ^2\mathrm{cot}\vartheta ,$$
(4.7)
whose eigenfunctions are the spin-weighted spherical harmonics $`{}_{๐ฌ}{}^{}Y_{jm}^{}`$ ,
$$\widehat{๐}[๐ฌ,\vartheta ,\phi ]_๐ฌY_{jm}(\vartheta ,\varphi )=(๐ฌ^2j(j+1))_๐ฌY_{jm}(\vartheta ,\varphi ).$$
(4.8)
Here $`j`$ and $`m`$ are such that $`j|๐ฌ|`$ and $`jmj`$. One has $`{}_{๐ฌ}{}^{}Y_{jm}^{}=0`$ for $`j<|๐ฌ|`$. \[Notice that we use the bold-faced $`๐ฌ`$ for the spin weight.\] The following relations between harmonics with different values of the spin weight $`๐ฌ`$ are useful:
$$\widehat{}^\pm [๐ฌ,\vartheta ,\phi ]_๐ฌY_{jm}=\pm \sqrt{(j\pm ๐ฌ)(j๐ฌ+1)}_{๐ฌ1}Y_{jm},$$
(4.9)
where
$$\widehat{}^\pm [๐ฌ,\vartheta ,\phi ]=\frac{}{\vartheta }\frac{i}{\mathrm{sin}\vartheta }\frac{}{\varphi }\pm ๐ฌ\mathrm{cot}\vartheta .$$
(4.10)
The harmonics for a fixed $`๐ฌ`$ form an orthonormal set on $`S^2`$.
Using the above definitions, Eqs. (4.4) can be represented as
$$\left(\widehat{๐}[๐ฌ_{ab}^1,\vartheta _1,\phi _1]+\widehat{๐}[๐ฌ_{ab}^2,\vartheta _2,\phi _2](๐ฌ_{ab}^1)^2(๐ฌ_{ab}^2)^2+2+\epsilon \right)\mathrm{\Phi }_{ab}=0,$$
(4.11)
where $`1a,b4`$ (no summation over $`a,b`$). Here the nonzero elements of the symmetric matrices $`๐ฌ_{ab}^1`$ and $`๐ฌ_{ab}^2`$ are
$`๐ฌ_{11}^1=๐ฌ_{22}^1=2,๐ฌ_{13}^1=๐ฌ_{14}^1=๐ฌ_{23}^1=๐ฌ_{24}^1=1,`$
$`๐ฌ_{33}^2=๐ฌ_{44}^2=2,๐ฌ_{13}^2=๐ฌ_{23}^2=๐ฌ_{14}^2=๐ฌ_{24}^2=1.`$ (4.12)
The solution to Eqs. (4.11) reads
$$\mathrm{\Phi }_{ab}=C_{ab}{}_{๐ฌ_{ab}^1}{}^{}Y_{j_1m_1}^{}(\vartheta _1,\phi _1)_{๐ฌ_{ab}^2}Y_{j_2m_2}(\vartheta _2,\phi _2),$$
(4.13)
with $`C_{ab}`$ being integration constants. The eigenvalue, $`\epsilon `$, is the same for all $`\mathrm{\Phi }_{ab}`$:
$$\epsilon =j_1(j_1+1)+j_2(j_2+1)2.$$
(4.14)
This is essentially the sum of squares of the two SO(3) angular momentum operators acting independently on the two spheres.
Let us now count the degeneracy of the modes. For this one should take into account the additional conditions (4.5), which gives algebraic constraints for the coefficients $`C_{ab}`$. We consider first the trace condition $`\varphi _\mu ^\mu =0`$. In the language of the tetrad projections this reduces to $`\mathrm{\Phi }_{12}+\mathrm{\Phi }_{34}=0`$, or equivalently to
$$C_{12}+C_{34}=0.$$
(4.15)
Hence only 9 out of the 10 constants $`C_{ab}`$ are independent.
Next, we consider the Lorenz condition $`_\sigma \varphi _\mu ^\sigma =0`$. This implies
$`\widehat{}^{}[๐ฌ_{a1}^1,\vartheta _1,\phi _1]\mathrm{\Phi }_{a1}+\widehat{}^+[๐ฌ_{a2}^1,\vartheta _1,\phi _1]\mathrm{\Phi }_{a2}`$ (4.16)
$`+`$ $`\widehat{}^{}[๐ฌ_{a3}^2,\vartheta _2,\phi _2]\mathrm{\Phi }_{a3}+\widehat{}^+[๐ฌ_{a4}^2,\vartheta _2,\phi _2]\mathrm{\Phi }_{a4}=0`$
(no summation over $`a`$). Inserting the solution (4.13) and using the recurrence relations in (4.9), these conditions reduce to algebraic constraints
$`\kappa _1C_{11}\alpha _1C_{12}+\alpha _2(C_{13}C_{14})=0,`$
$`\alpha _1C_{12}\kappa _1C_{22}+\alpha _2(C_{23}C_{24})=0,`$
$`\alpha _1(C_{13}C_{23})+\kappa _2C_{33}\alpha _2C_{34}=0,`$
$`\alpha _1(C_{14}C_{24})+\alpha _2C_{34}\kappa _2C_{44}=0.`$ (4.17)
Here $`\alpha _\iota =\sqrt{j_\iota (j_\iota +1)}`$ (with $`\iota =1,2`$) and $`\kappa _\iota =\sqrt{(j_\iota +2)(j_\iota 1)}`$ for $`j_\iota 1`$ while $`\kappa _\iota =0`$ for $`j_\iota =0`$.
For $`j_\iota 2`$ (which corresponds to quadrupole or higher deformations of each sphere) none of the coefficients $`\alpha _\iota `$, $`\kappa _\iota `$ vanish, and the algebraic constraints (4.17) reduce the number of independent coefficients $`C_{ab}`$ to five. This gives the degeneracy $`d`$:
$$j_12,j_22,d=5(2j_1+1)(2j_2+1).$$
(4.18)
The situation is different for small values of $`j_\iota `$. Consider, for example, the $`j_1=j_2=0`$ sector. Since the harmonics $`{}_{๐ฌ}{}^{}Y_{jm}^{}`$ vanish for $`j<|๐ฌ|`$, we must set in the solution (4.13) all $`C_{ab}`$โs to zero, apart from $`C_{12}=C_{34}`$. The Lorenz condition (4.16) is then fulfilled. As a result, there is only one independent integration constant, which yields
$$j_1=j_2=0,d=1.$$
(4.19)
Notice that in this case $`\epsilon =2`$.
In a similar way one can consider the sector where $`j_1=0`$ and $`j_2=1`$ (or $`j_1=1`$ and $`j_2=0`$), in which case $`\epsilon =0`$. One discovers then that the Lorenz constraints (4.16) require that all non-trivial coefficients $`C_{ab}`$ mush vanish. As a result,
$$j_1=0,j_2=1,\mathrm{or}j_2=1,j_1=0,d=0,$$
(4.20)
which shows that there are no zero modes.
Next,
$$j_1=j_2=1,d=(2j_1+1)(2j_2+1)=9;$$
(4.21)
and finally
$`j_12,j_2=0,d=2j_1+1;`$
$`j_12,j_2=1,d=9(2j_1+1),`$ (4.22)
which conditions are symmetric under interchanging $`j_1`$ and $`j_2`$.
To recapitulate, the spectrum of the tensor fluctuations contains one negative mode and no zero modes.
#### 4.1.2 Vector modes
Let us now consider the eigenvalue problem
$$(_\alpha ^\alpha \mathrm{\Lambda })\eta _\mu =\sigma \eta _\mu $$
(4.23)
subject to the condition
$$_\mu \eta ^\mu =0$$
(4.24)
for the vector fluctuations on the $`S^2\times S^2`$ background. We again expand the fluctuations with respect to the basis (4.1),
$$\eta _\mu =e_\mu ^a\mathrm{\Psi }_a,$$
(4.25)
insert this into (4.23), and project back to the tetrad. Similarly as in the tensor case, the equations decouple to give
$$(\widehat{๐}[๐ฌ_a^1,\vartheta _1,\phi _1]+\widehat{๐}[๐ฌ_a^2,\vartheta _2,\phi _2]+1+\sigma )\mathrm{\Psi }_a=0,$$
(4.26)
where $`1a4`$ (no summation over $`a`$), and nonzero coefficients read $`๐ฌ_1^1=๐ฌ_2^1=๐ฌ_3^2=๐ฌ_4^2=1`$. The solution is
$$\mathrm{\Psi }_a=C_a{}_{๐ฌ_a^1}{}^{}Y_{j_1m_1}^{}(\vartheta _1,\phi _1)_{๐ฌ_a^2}Y_{j_2m_2}(\vartheta _2,\phi _2),$$
(4.27)
with $`C_a`$ being integration constants, and the eigenvalue is the same for all $`\mathrm{\Psi }_a`$:
$$\sigma =j_1(j_1+1)+j_2(j_2+1)2.$$
(4.28)
The Lorenz condition, $`_\sigma \eta ^\sigma =0`$, reads
$`\widehat{}^{}[๐ฌ_1^1,\vartheta _1,\phi _1]\mathrm{\Psi }_1+\widehat{}^+[๐ฌ_2^1,\vartheta _1,\phi _1]\mathrm{\Psi }_2`$ (4.29)
$`+`$ $`\widehat{}^{}[๐ฌ_3^2,\vartheta _2,\phi _2]\mathrm{\Psi }_3+\widehat{}^+[๐ฌ_4^2,\vartheta _2,\phi _2]\mathrm{\Psi }_4=0,`$
which reduces upon inserting (4.27) to the algebraic condition
$$\sqrt{j_1(j_1+1)}(C_1C_2)+\sqrt{j_2(j_2+1)}(C_3C_4)=0.$$
(4.30)
This allows one to count the degeneracies:
$$j_11,j_21(\sigma >0),d=3(2j_1+1)(2j_2+1);$$
(4.31)
and also
$`j_12,j_2=0(\sigma >0),d=j_1(j_1+1);`$
$`j_1=1,j_2=0(\sigma =0),d=3;`$
$`j_1=0,j_2=0(\sigma =2),d=0.`$ (4.32)
These results are symmetric under $`j_1j_2`$, hence there are no negative modes, there are six zero modes corresponding to the six Killing vectors of $`S^2\times S^2`$, and the rest of the spectrum is positive.
#### 4.1.3 Scalar modes and the orthogonality conditions
The eigenvalue problem for the scalar modes,
$$_\alpha ^\alpha h=\lambda h,$$
(4.33)
reduces to the equation
$$(\widehat{๐}[0,\vartheta _1,\phi _1]+\widehat{๐}[0,\vartheta _2,\phi _2]+\lambda )h=0,$$
(4.34)
whose solutions are
$$h=Y_{j_1m_1}(\vartheta _1,\phi _1)Y_{j_2m_2}(\vartheta _2,\phi _2)$$
(4.35)
(for $`๐ฌ=0`$ the spin-weighted spherical harmonics coincide with the usual spherical harmonics). The eigenvalues are just
$$\lambda =j_1(j_1+1)+j_2(j_2+1),$$
(4.36)
and the degeneracies are
$$j_10,j_20,d=(2j_1+1)(2j_2+1).$$
(4.37)
We have obtained the spectra of all relevant fluctuation operators. Although the eigenfunctions are complex, one can pick up their real part in a way that is consistent with the orthogonality conditions. For example, for the scalar modes one considers
$$\mathrm{}(h)=\frac{1+i}{\sqrt{2}}Y_{j_1m_1}Y_{j_2m_2}+c.c,$$
(4.38)
and one can easily see that the modes $`\mathrm{}(h)`$ with different quantum numbers $`(j_1m_1j_2m_2)`$ are orthogonal with respect to the scalar product defined in Eq. (3.11).
For the vector modes $`\mathrm{\Psi }_a`$ the procedure is slightly more complicated, since the tetrad metric $`\eta _{ab}`$ is not diagonal. In addition, harmonics $`{}_{๐ฌ}{}^{}Y_{jm}^{}`$ for different values of the spin weight are not orthogonal. Consider, however, the real combinations
$$\mathrm{}(\eta _\mu )=\frac{1+i}{\sqrt{2}}e_\mu ^a\mathrm{\Psi }_a+c.c,$$
(4.39)
where $`\mathrm{\Psi }_a`$ has quantum numbers $`(j_1m_1j_2m_2)`$. Consider $`\mathrm{}(\eta _\mu ^{(1)})`$ and $`\mathrm{}(\eta _\mu ^{(2)})`$ with different quantum numbers. Their scalar product (defined in Eq. (3.11)) can be computed using the relations
$$\eta _{ab}=e_a^\mu e_b^\nu g_{\mu \nu },\delta _{ab}=e_a^\mu (e_b^\nu )^{}g_{\mu \nu },$$
(4.40)
which gives
$`\mathrm{}(\eta _\mu ^{(1)}),\mathrm{}(\eta ^{(2)\mu })={\displaystyle \underset{a}{}}\mathrm{\Psi }_a^{(1)},\mathrm{\Psi }_a^{(2)}`$ (4.41)
$`+i\mathrm{\Psi }_1^{(1)},\mathrm{\Psi }_2^{(2)}+i\mathrm{\Psi }_3^{(1)},\mathrm{\Psi }_4^{(2)}+c.c.`$
Here each term in the sum $`_a\mathrm{\Psi }_a^{(1)},\mathrm{\Psi }_a^{(2)}`$ is a bilinear combination of spin-weighted harmonics with the same value of the spin weight, such that the orthogonality relation holds. Next, integrating by parts and using the recurrence relations (4.9) one can show that the remaining term in the scalar product, $`i\mathrm{\Psi }_1^{(1)},\mathrm{\Psi }_2^{(2)}+i\mathrm{\Psi }_3^{(1)},\mathrm{\Psi }_4^{(2)}+c.c`$, vanishes. This shows that vector modes $`\mathrm{}(\eta _\mu )`$ with different quantum numbers are orthogonal.
A similar procedure can be carried out for the tensor modes. Hence for all eigenmodes considered above one can choose the real part in such a way that the orthogonality condition holds. This is a manifestation of the fact that the fluctuation operators are self-adjoint. We finally restore the dependence on $`\mathrm{\Lambda }`$ and summarize the results of this section in Tab.1. There is one negative mode in the spectrum, and this plays a crucial role in our analysis. The corresponding deformation increases the radius of one of the spheres, shrinking at the same time the second one.
### 4.2 Fluctuations around the $`S^4`$ instanton
The $`S^4`$ instanton can be viewed as the four-dimensional sphere with radius $`\sqrt{3/\mathrm{\Lambda }}`$ in five-dimensional Euclidean space $`E^5`$. Although the corresponding eigenvalue problem for fluctuations was considered in , we have reanalyzed it for the sake of completeness (with the same result) and shall present below the key steps of our analysis. The problem essentially reduces to studying representations of SO(5) , whose Casimir operator can be related to the invariant Laplacians on $`S^4`$ with the help of the projection formalism . We shall therefore first outline the group theory part by summarizing the relevant facts about representations of $`SO(5)`$. We shall work on the unit 4-sphere rescaling at the end the eigenvalues by the factor $`\mathrm{\Lambda }/3`$.
#### 4.2.1 Representations of SO(5)
The unit sphere $`S^4`$ in $`E^5`$ is defined in Cartesian coordinates by the equation $`_{a=1}^5(x^a)^2=1`$. It is convenient to use the complex coordinates $`\xi ^{\pm 1}=(x^1\pm ix^2)/\sqrt{2}`$, $`\xi ^{\pm 2}=(x^3\pm ix^4)/\sqrt{2}`$, $`\xi ^0=x^5`$. We shall not distinguish between lower and upper indices, $`\xi _i=\xi ^i`$. In these new coordinates the defining quadratic form reads $`_{i=2}^2\xi ^i\xi ^i=1`$, which is annihilated by the generators of SO(5):
$$Y_j^i=\xi ^i\frac{}{\xi ^j}\xi ^j\frac{}{\xi ^i},$$
(4.42)
whose commutation relations are
$$[Y_j^i,Y_l^k]=\delta _j^kY_l^i\delta _l^iY_j^k+\delta _j^lY_k^i\delta _i^kY_l^jC_{ij,kl}^{pq}Y_q^p.$$
(4.43)
Since $`Y_j^i=Y_i^j`$, the independent generators can be chosen to be those for $`i<j`$. $`Y_1^1`$ and $`Y_2^2`$ generate the Cartan subalgebra, while $`Y_j^i`$ and $`Y_i^j`$ for $`i<j<i`$ are the raising and lowering operators, respectively. One has
$$[Y_i^i,Y_l^k]=\alpha _l^k(i)Y_l^k,$$
(4.44)
where
$$\alpha _l^k(i)=\delta _k^i\delta _l^i+\delta _l^i\delta _k^i$$
(4.45)
determine the root vectors with the components $`\alpha _l^k(\alpha _l^k(1),\alpha _l^k(2))`$. The roots corresponding to the four raising operators are $`\alpha _1^2=(1,1)`$, $`\alpha _0^2=(0,1)`$, $`\alpha _1^2=(1,1)`$, and $`\alpha _0^1=(1,0)`$.
Irreducible representations of SO(5) are characterized by two numbers denoted by $`m(m_1,m_2)`$, where $`m_2m_1`$ and both $`m_1`$ and $`m_2`$ are either integer or half-integer. The highest weight vector $`\psi _m`$ is annihilated by all raising operators, $`Y_j^i\psi _m=0`$ for $`i>j>i`$, and it is an eigenvector of the Cartan subalgebra generators, $`Y_i^i\psi _m=m_i\psi _m`$, $`i=1,2`$. Using these properties and also $`[Y_j^i,Y_i^j]=Y_i^iY_j^j2\delta _j^iY_j^i`$, one finds the eigenvalues of the Casimir operator,
$$\widehat{C}\psi _m\frac{1}{2}\underset{i,j}{}Y_j^iY_i^j\psi _m=C_m\psi _m,$$
(4.46)
where
$$C_m=m_1(m_1+1)+m_2(m_2+3)$$
(4.47)
is the same for all vectors of the representation. The dimension of representations is given by
$$\mathrm{dim}(m)=\underset{\alpha }{}\alpha ,r+m/\underset{\alpha }{}\alpha ,r.$$
(4.48)
Here the product is over the four root vectors described above, and $`r=\frac{1}{2}_\alpha \alpha =(\frac{1}{2},\frac{3}{2})`$. One has $`r+m=(\frac{1}{2}+m_1,\frac{3}{2}+m_2)`$, and $`,`$ is the scalar with respect to the Cartan metric $`g_{ij}=C_{ii,kl}^{pq}C_{jj,pq}^{kl}=6\delta _{ij}`$ (here $`i,j=1,2`$). As a result,
$$\mathrm{dim}(m)=\frac{1}{6}(2m_1+1)(2m_2+3)(m_2m_1+1)(m_1+m_2+2).$$
(4.49)
#### 4.2.2 Scalar modes
Using Eqs.(4.47),(4.49) one can find the spectra of the relevant fluctuation operators. It is now convenient to pass back to the Cartesian coordinates $`x^a=x_a`$ ($`a=1,\mathrm{}5`$), such that the sphere $`S^4`$ is determined by the condition $`r\sqrt{x^ax_a}=1`$. Let $`n^ax^a/r`$ be the unit normal to the sphere. The (anti-hermitean) generators of SO(5) in Cartesian coordinates are given by $`M_{ab}=n_a_bn_b_a`$, and the Casimir operator is $`\widehat{C}=\frac{1}{2}(M_{ab})^2\frac{1}{2}_{ab}(M_{ab})^2`$.
Let us define the projection operator $`P_{ab}=\delta _{ab}n_an_b=P^{ab}=P_b^a`$, which can be thought of as the induced metric on the sphere. In what follows we shall use the projection formalism to describe geometrical 4-objects tangent to the sphere in terms of 5-objects of the embedding space. For example, a 4-vector $`\eta _\mu `$ can be described as a 5-vector $`\eta _a`$ subject to the condition $`n^a\eta _a=0`$. The covariant derivative of a tensor is obtained by taking the partial derivative and then projecting all the indices down to the sphere. For example, $`_a\eta _b=(_p\eta _q)P_a^pP_b^q`$. One has $`n_a=n^a`$, while for objects tangent to the sphere 5-indices can be raised and lowered either with $`P_{ab}`$ or with $`\delta _{ab}`$. The curvature tensor is given by $`R_{psqt}=P_{pq}P_{st}P_{pt}P_{sq}`$.
Consider first scalar fluctuations. The covariant Laplacian for a scalar field $`h`$ can be expressed in terms of the angular momentum operator as
$$\mathrm{}hP^{ab}_a(P_b^c_ch)=\frac{1}{2}(M_{ab})^2h=\widehat{C}h.$$
(4.50)
Scalars transform according to the $`(0,j)`$ representations, which correspond to the Young tableaux โฆ and can be represented in terms of homogeneous polynomials on $`S^4`$ as
$$h=h_{(a_1\mathrm{}a_j)}n^{a_1}\mathrm{}n^{a_j}.$$
(4.51)
Hence, the eigenvalues of the Casimir operator in Eq. (4.50) are $`C_{(0,j)}=j(j+3)`$, which gives the spectrum of the scalar eigenvalue problem, $`\mathrm{\Delta }_0h=\lambda h`$ with $`\mathrm{\Delta }_0\frac{\mathrm{\Lambda }}{3}\mathrm{}`$:
$$\lambda =\frac{\mathrm{\Lambda }}{3}j(j+3),d=\frac{1}{6}(2j+3)(j+2)(j+1),j0.$$
(4.52)
#### 4.2.3 Vector modes
Consider a tangent vector field $`\eta _s=P_s^a\eta _a`$. The invariant Laplacian reads
$$\mathrm{}\eta _sP^{ab}_a(P_b^c_c\eta _pP_q^p)P_s^q=\frac{1}{2}(M_{ab})^2\eta _s+2(_a\eta ^a)n_s+\eta _s.$$
(4.53)
Here the last two terms on the right can be related to the contribution of the spin operator. Under general SO(5) rotations a vector $`\eta (x)`$ transforms into $`\stackrel{~}{\eta }(x)=R\eta (R^1x)`$, where $`R=\mathrm{exp}(\omega ^{ab}S_{ab})`$ with $`\omega ^{ab}=\omega ^{ba}`$ being the rotation parameters and $`S_{ab}(S_{ab})^{pq}=\delta _a^p\delta _b^q\delta _b^p\delta _a^q`$. For $`|\omega _{ab}|1`$ one obtains $`\stackrel{~}{\eta }\eta =\omega ^{ab}(M_{ab}+S_{ab})\eta `$, such that $`S_{ab}`$ can be identified with the spin operator: $`(S_{ab}\eta )_s=(S_{ab})_s^p\eta _p`$. As a result,
$$\mathrm{}\eta _s=\left(\frac{1}{2}(M_{ab}+S_{ab})^2+3\right)\eta _s(\widehat{C}+3)\eta _s,$$
(4.54)
where the Casimir operator is now the square of the total angular momentum. The general vector harmonics on $`S^4`$ can be obtained by considering the product of a constant vector in $`E^5`$ with scalar harmonics on $`S^4`$. Such a product decomposes into three irreducible pieces, $`(0,1)(0,j)=(1,j)(0,j+1)(0,j1)`$, which can be visualized as
$$\text{}\text{}\text{ }\text{ โฆ}\text{ }=\text{}\text{ }\text{ โฆ}\text{ }\text{}\text{ }\text{ โฆ}\text{ }\text{}\text{ }\text{ โฆ}\text{ }.$$
(4.55)
The first term on the right here is the ($`1,j`$)-piece, and in the language of homogeneous polynomials it reads
$$\eta _s=\eta _{[sa](a_1\mathrm{}a_{j1})}n^an^{a_1}\mathrm{}n^{a_{j2}}n^{a_{j1}},$$
(4.56)
where $`\eta _{[sa](a_1\mathrm{}a_{j1})}`$ is traceless with respect to any pair of indices. This is manifestly tangential and coexact. As a result, the eigenvalues of the Casimir operator are $`C_{(1,j)}=j(j+3)+2`$, and the spectrum of the vector eigenvalue problem $`\mathrm{\Delta }_1\eta _s(\frac{\mathrm{\Lambda }}{3}\mathrm{}\mathrm{\Lambda })\eta _s=\sigma \eta _s`$ in the sector where $`_a\eta ^a=n^a\eta _a=0`$ is given by
$$\sigma =\frac{\mathrm{\Lambda }}{3}(j(j+3)4),d=\frac{1}{2}j(j+3)(2j+3),j1.$$
(4.57)
One can also directly verify that the harmonic $`\eta _s`$ in Eq.(4.56) fulfills the condition $`\frac{1}{2}(M_{ab})^2\eta _s=j(j+3)\eta _s`$. It follows then from Eq. (4.53) that $`\mathrm{}\eta _s=(j(j+3)1)\eta _s`$, and this again yields the spectrum in Eq. (4.57). The correct degeneracy can be obtained by counting the independent components of $`\eta _{[sa](a_1\mathrm{}a_{j1})}`$ .
The remaining two pieces in (4.55), when represented in terms of the polynomials on $`S^4`$, can be related to the exact tangential and the normal components of the vector field.
#### 4.2.4 Tensor modes
For a symmetric tensor $`h_{pq}=P_p^aP_q^bh_{ab}`$ a direct calculation gives
$`\mathrm{}h_{pq}+2R_{psqt}h^{st}`$ $``$ $`P^{ab}_a(P_b^c(_ch_{\overline{p}\overline{q}})P_{\underset{ยฏ}{p}}^{\overline{p}}P_{\underset{ยฏ}{q}}^{\overline{q}})P_p^{\underset{ยฏ}{p}}P_q^{\underset{ยฏ}{q}}+2(P_{pq}P_{st}P_{pt}P_{sq})h^{st}`$ (4.58)
$`=`$ $`{\displaystyle \frac{1}{2}}(M_{ab})^2h_{pq}+2n_p(^ah_{aq})+2n_q(^ah_{ap})+2\delta _{pq}h_a^a`$
$`=`$ $`\left({\displaystyle \frac{1}{2}}(M_{ab}+\mathrm{\Sigma }_{ab})^2+6\right)h_{pq}(\widehat{C}+6)h_{pq}.`$
Here the spin operator is defined in the same way as for vectors, which gives $`(\mathrm{\Sigma }_{ab}h)_{pq}=(S_{ab})_p^sh_{sq}+(S_{ab})_q^sh_{sp}`$. The general tensor harmonics on $`S^4`$ are obtained by the direct products $`(0,2)(0,j)=(0,j+2)(1,j+1)(0,j)(2,j)(0,j+1)(1,j1)(0,j2)`$. Again this can be visualized by Youngโs diagrams and represented in the language of homogeneous polynomials. The transverse and tracefree harmonics tangent to the sphere correspond to the $`(2,j)`$ piece, whose explicit representation is
$$h_{pq}=h_{[pa][qb](a_1\mathrm{}a_{j2})}n^an^bn^{a_1}\mathrm{}n^{a_{j2}}.$$
(4.59)
Here $`h_{[pa][qb](a_1\mathrm{}a_{j2})}`$ is traceless with respect to any pair of indices and is symmetric under interchange of the $`[pa]`$ and $`[qb]`$ pairs. As a result, the eigenvalues of the Casimir operator are $`C_{(2,j)}=j(j+3)+6`$. This gives the spectrum of the tensor eigenvalue problem $`\mathrm{\Delta }_2h_{pq}=\epsilon h_{pq}`$ in the sector where $`^ah_{ab}=n^ah_{ab}=h_a^a=0`$:
$$\epsilon =\frac{\mathrm{\Lambda }}{3}j(j+3),d=\frac{5}{6}(j1)(j+4)(2j+3),j2.$$
(4.60)
The same result can be obtained by directly verifying that $`h_{pq}`$ in Eq. (4.59) fulfills the condition $`\frac{1}{2}(M_{ab})^2h_{pq}=j(j+3)h_{pq}`$.
The other tensor harmonics in the expansion of $`(0,2)(0,j)`$ correspond to the exact and coexact pieces of the longitudinal vector part of the 4-metric, to those of the 4-vector $`h_{5\mu }`$, and to the trace .
We summarize the results of our analysis in this section in Tab.2. Notice that the scalar and tensor eigenvalues are the same (for $`j2`$), while the vector spectrum is shifted by a constant.
## 5 Partition function
Now we are able to derive the explicit expressions for the one-loop partition functions for fluctuations around the $`S^2\times S^2`$ and $`S^4`$ instantons. The corresponding formula was obtained in Eq.(3.50) above. It is convenient to pass from $`\mu _o`$ to the dimensionless normalization parameter $`\mu _0`$ via the rescaling
$$\mu _o=\sqrt{\mathrm{\Lambda }}\mu _0.$$
(5.1)
The one-loop partition function for gravity fluctuations around an Euclidean background then reads
$$Z[g_{\mu \nu }]=\frac{\mu _0}{\sqrt{2}}\mathrm{\Omega }_0\frac{\mathrm{\Omega }_2}{\mathrm{\Omega }_1}\sqrt{\frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}\mathrm{e}^I.$$
(5.2)
Here the first two factors on the right are the contributions of the scalar modes. The factor $`\mu _0/\sqrt{2}`$ is due to the constant conformal mode, which is always present, and $`\mathrm{\Omega }_0`$ is the contribution of the 5 scalar modes with eigenvalue $`4\mathrm{\Lambda }/3`$ which exist only for the $`S^4`$ instanton (see Tab.2):
$$\mathrm{\Omega }_0=\left(\sqrt{\frac{2}{3}}\frac{1}{\mu _0}\right)^5.$$
(5.3)
For any background other than $`S^4`$ one has $`\mathrm{\Omega }_0=1`$. As was discussed above, other scalar modes do not contribute to the partition function.
The factor $`\mathrm{\Omega }_2`$ in (5.2) is the contribution of the negative tensor mode,
$$\mathrm{\Omega }_2=\frac{\mu _0}{2i\sqrt{|\epsilon _{}|}}.$$
(5.4)
For the $`S^2\times S^2`$ instanton there is one such mode with $`\epsilon =2`$, while in the $`S^4`$ case all tensor modes are positive and $`\mathrm{\Omega }_2=1`$. Next,
$$\mathrm{\Omega }_1=\left(\underset{j}{}\frac{\mu _0^2}{\sqrt{\pi }}\mathrm{\Lambda }\frac{}{C_j}\right)Vol(),$$
(5.5)
is the isometry factor. If the background has no isometries then $`\mathrm{\Omega }_1=1`$.
The determinants in Eq. (5.2) are the contributions of the positive vector and tensor modes. One has
$$\sqrt{\mathrm{Det}^{}\mathrm{\Delta }_1}=\left(\underset{s}{\overset{}{}}\sqrt{\frac{\sigma _s}{\mathrm{\Lambda }\mu _0^2}}\right)=\mathrm{exp}\left(\frac{1}{2}\zeta _1^{}(0)\frac{1}{2}(\mathrm{ln}\mu _0^2)\zeta _1(0)\right),$$
(5.6)
where the $`\zeta `$-function for the positive, transverse vector modes is
$$\zeta _1(z)=\underset{s}{\overset{}{}}\left(\frac{\mathrm{\Lambda }}{\sigma _s}\right)^z.$$
(5.7)
Similarly for the positive, transverse traceless tensor modes:
$$\sqrt{\mathrm{Det}^{}\mathrm{\Delta }_2}=\left(\underset{k}{\overset{}{}}\sqrt{\frac{ฯต_k}{\mathrm{\Lambda }\mu _0^2}}\right)=\mathrm{exp}\left(\frac{1}{2}\zeta _2^{}(0)\frac{1}{2}(\mathrm{ln}\mu _0^2)\zeta _2(0)\right)$$
(5.8)
with
$$\zeta _2(z)=\underset{k}{\overset{}{}}\left(\frac{\mathrm{\Lambda }}{ฯต_k}\right)^z.$$
(5.9)
The last factor in Eq. (5.2) is the classical contribution, with $`I`$ being the action for the background. Let us now apply these formulas.
### 5.1 The $`S^2\times S^2`$ instanton
The classical action is $`I[S^2\times S^2]=2\pi /\mathrm{\Lambda }G`$, and according to Tab.1,
$$\mathrm{\Omega }_0=1,\mathrm{\Omega }_2=\frac{\mu _0}{2i\sqrt{2}}.$$
(5.10)
Consider now the isometry factor $`\mathrm{\Omega }_1`$ in (5.5), which is due to the background SO(3)$`\times `$SO(3) symmetry. Each of the two SO(3) groups can be parameterized by matrices $`U_{ik}=\mathrm{exp}(\epsilon _{ikj}C_j`$). The invariant metric on the SO(3) space is $`g_{ik}=\frac{1}{2}\mathrm{tr}(_iU_kU^1)\delta _{ik}`$ for $`C_j0`$. The Haar measure is $`d\mu (C)=\sqrt{\mathrm{det}g_{ik}}dC_1dC_2dC_3`$, and the volume $`Vol`$(SO(3))= $`๐\mu (C)=8\pi ^2`$. For later use, we reproduce this result in a different way. The measure for a (compact, semi-simple) Lie group $`๐ข`$ can be represented as the product of the measure for the maximal subgroup $``$ and that for the coset $`๐ข/`$. This implies that
$$Vol(๐ข)=Vol()\times Vol(๐ข/).$$
(5.11)
In particular, $`Vol`$(SO(3))=$`Vol`$(SO(2))$`\times Vol(S^2)`$, where $`Vol`$(SO(2))= $`2\pi `$, and the volume of the $`S^2`$ coset with unit (due to the normalization of the measure) radius is $`Vol(S^2)=4\pi `$. As a result, $`Vol`$(SO(3))=$`2\pi \times 4\pi =8\pi ^2`$.
When acting on $`S^2`$, the SO(3) generators $`\frac{}{C_j}`$ generate rotations in the three orthogonal planes of the embedding Euclidean 3-space. Let $`\frac{}{C_3}`$ be the generator of rotations in the XY-plane, such that the azimuthal angle of the spherical coordinate system changes as $`\phi \phi +C_3`$. Then the norm $`\frac{}{C_3}`$ is the square root of
$$\frac{}{\phi },\frac{}{\phi }=\frac{1}{32\pi G}_{S^2\times S^2}g_{\phi \phi }\sqrt{g}d^4x=\frac{\pi }{3\mathrm{\Lambda }^3G}.$$
(5.12)
Obviously, the norms $`\frac{}{C_1}`$ and $`\frac{}{C_2}`$ and those of the generators of the second SO(3) factor are the same. Hence,
$$\mathrm{\Omega }_1=\left(\frac{\mu _0^2}{\sqrt{\pi }}\mathrm{\Lambda }\frac{}{\phi }\right)^6\left(Vol(\mathrm{SO}(3))\right)^2=\frac{64\pi ^4(\mu _0)^{12}}{27(\mathrm{\Lambda }G)^3}.$$
(5.13)
Consider now the positive modes. The $`\zeta `$-function associated with the positive vector modes is (see Tab.1)
$$\zeta _1(s)=\underset{j=2}{\overset{\mathrm{}}{}}\frac{2(2j+1)}{\{j(j+1)2\}^s}+\underset{j_1=2}{\overset{\mathrm{}}{}}\underset{j_2=2}{\overset{\mathrm{}}{}}\frac{3(2j_1+1)(2j_2+1)}{\{j_1(j_1+1)+j_2(j_2+1)2\}^s}.$$
(5.14)
This can be represented as
$$\zeta _1(s)=4^s(2\zeta (2,9|s)+3Z(1,10|s)),$$
(5.15)
where the following two functions have been introduced:
$`\zeta (k,\nu |s)`$ $`=`$ $`{\displaystyle \underset{j=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2j+1}{\{(2j+1)^2+\nu \}^s}},`$ (5.16)
$`Z(k,\nu |s)`$ $`=`$ $`{\displaystyle \underset{j_1=k}{\overset{\mathrm{}}{}}}{\displaystyle \underset{j_2=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(2j_1+1)(2j_2+1)}{\left\{(2j_1+1)^2+(2j_2+1)^2+\nu \right\}^s}}.`$ (5.17)
These functions are studied in detail in the Appendix. Similarly, using the results of Tab.1 one obtains the $`\zeta `$-function for the positive tensor modes
$$\zeta _2(s)=9\times 2^s+4^s(2\zeta (2,9|s)+18\zeta (2,1|s)+5Z(2,10|s)).$$
(5.18)
The following relation implied by the definitions in (5.16), (5.17), will be useful: $`Z(1,10|s)=Z(2,10|s)+6\zeta (2,1|s)+9\times 8^s`$.
#### 5.1.1 The scaling behaviour
Before we proceed further, it is very instructive to pause and check whether the expressions above agree with the general formulas for the scaling behaviour of effective actions. We shall follow the approach of Christensen and Duff , who relate this scaling behaviour to
$`N_0`$ $`=`$ $`{\displaystyle \frac{1}{180(4\pi )^2}}{\displaystyle (R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }+636\mathrm{\Lambda }^2)\sqrt{g}d^4x},`$
$`N_1`$ $`=`$ $`{\displaystyle \frac{1}{180(4\pi )^2}}{\displaystyle (11R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }+984\mathrm{\Lambda }^2)\sqrt{g}d^4x},`$
$`N_2`$ $`=`$ $`{\displaystyle \frac{1}{180(4\pi )^2}}{\displaystyle (189R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }756\mathrm{\Lambda }^2)\sqrt{g}d^4x}.`$ (5.19)
Here $`N_0`$ is the โnumber of eigenvaluesโ of the scalar operator $`\mathrm{\Delta }_02\mathrm{\Lambda }`$ acting on a manifold with $`R_{\mu \nu }=\mathrm{\Lambda }g_{\mu \nu }`$. $`N_1`$ is the number of eigenvalues of the vector operator $`\mathrm{\Delta }_1`$ acting in the space of all vectors, that is, including both transverse and longitudinal fluctuations. Finally, $`N_2`$ counts both transverse and longitudinal eigenstates of the tensor operator $`\mathrm{\Delta }_2`$, with the only requirement that the fluctuations must be traceless.
Let us apply these formulas to the $`S^2\times S^2`$ background. The volume of the manifold is $`V_{S^2\times S^2}=(4\pi )^2/\mathrm{\Lambda }^2`$, while $`R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }=8\mathrm{\Lambda }^2`$. As a result,
$$N_0=\frac{161}{45},N_1=\frac{224}{45},N_2=\frac{21}{5}.$$
(5.20)
Now let us obtain the same result via a direct evaluation of the $`\zeta `$-functions. First we consider the scalar case. Using the results of Tab.1, the operator $`\mathrm{\Delta }_02\mathrm{\Lambda }`$ has one negative mode, six zero modes, while the rest of the spectrum is positive and gives rise to the $`\zeta `$-function
$$\zeta _0(s)=4^s(2\zeta (2,9|s)+Z(1,10|s)).$$
(5.21)
Hence the number of all eigenvalues is $`7+\zeta _0(0)`$. In order to compute $`\zeta _0(0)`$, we use the results of the Appendix, where the following values are obtained:
$`\zeta (k,\nu |0)`$ $`=`$ $`{\displaystyle \frac{1}{12}}{\displaystyle \frac{1}{4}}\nu k^2,`$ (5.22)
$`Z(k,\nu |0)`$ $`=`$ $`{\displaystyle \frac{1}{32}}\nu ^2{\displaystyle \frac{1}{24}}\nu +2k^4+({\displaystyle \frac{1}{2}}\nu {\displaystyle \frac{2}{3}})k^2+{\displaystyle \frac{13}{360}}.`$ (5.23)
This gives for the $`\zeta `$-functions in (5.15), (5.18), (5.21)
$$\zeta _0(0)=\frac{154}{45},\zeta _1(0)=\frac{18}{5},\zeta _2(0)=\frac{38}{9}.$$
(5.24)
Using these, the number of scalar eigenvalues is $`N_0=7\frac{154}{45}=\frac{161}{45}`$, which agrees with (5.20).
Next, the vector operator $`\mathrm{\Delta }_1`$ has 6 zero modes, such that the number of its eigenvalues in the transverse sector is $`6+\zeta _1(0)`$. Now, one should take into account also the longitudinal vectors, which are gradients of scalars. It is not difficult to see that if $`_\mu \chi `$ is an eigenvector of $`\mathrm{\Delta }_1`$, such that $`\mathrm{\Delta }_1_\mu \chi =\sigma _\mu \chi `$, then $`(\mathrm{\Delta }_02\mathrm{\Lambda })\chi =\sigma \chi `$. We see that the eigenfunctions of $`\mathrm{\Delta }_02\mathrm{\Lambda }`$ are in one-to-one correspondence with the longitudinal vectors. The number of the latter is therefore $`N_01`$, where the one is subtracted because the ground state scalar eigenfunction is constant, which vanishes upon differentiation. We therefore conclude that $`N_1=6+\zeta _1(0)+N_01=6\frac{18}{5}+\frac{161}{45}1=\frac{224}{45}`$, which also agrees with (5.20).
Finally, the number of traceless eigenvalues of $`\mathrm{\Delta }_2`$ is $`1+\zeta _2(0)`$ (here the one is the contribution of the negative mode) plus the number of longitudinal traceless tensor harmonics $`\varphi _{\mu \nu }^\mathrm{L}=_\mu \xi _\nu +_\nu \xi _\mu \frac{1}{2}g_{\mu \nu }_\rho \xi ^\rho `$.
Now, if $`\mathrm{\Delta }_1\xi _\mu =\sigma \xi _\nu `$ then for $`\varphi _{\mu \nu }^\mathrm{L}`$ associated with $`\xi _\mu `$ one has $`\mathrm{\Delta }_2\varphi _{\mu \nu }^\mathrm{L}=\sigma \varphi _{\mu \nu }^\mathrm{L}`$. Hence, the number of longitudinal tensors is determined by the number of vectors, which gives $`N_2=1+\zeta _2(0)+(N_16)`$. Here six is subtracted because the six Killing vectors do not contribute to the tensor spectrum, since for Killing vectors one has $`\varphi _{\mu \nu }^\mathrm{L}=0`$. We therefore obtain $`N_2=1+\frac{38}{9}+\frac{224}{45}6=\frac{21}{5}`$, which again is in perfect agreement with (5.20).
The overall scale dependence of the partition function is given by the factor $`(\mu _0)^{N_2+N_02N_1}`$. For the $`S^2\times S^2`$ instanton one has $`N_2+N_02N_1=\frac{98}{45}`$, and we shall shortly see that this agrees with our analysis.
#### 5.1.2 The partition function $`Z[S^2\times S^2]`$
It is now a simple task to insert the formulas above into the expression for the partition function. We obtain
$$\sqrt{\frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}=\mathrm{exp}\left(\zeta ^{}(0)+\mathrm{ln}\mu _0^2\zeta (0)\right),$$
(5.25)
where
$$\zeta (s)\frac{1}{2}(\zeta _2(s)\zeta _1(s))=9\times 2^s+4^sZ(2,10|s).$$
(5.26)
Using the values $`Z(2,10|0)=\frac{581}{45}`$ and $`Z^{}(2,10|0)\mathrm{{\rm Y}}=18.3118`$ (see Eq.(A.51) in the Appendix) we find
$$\sqrt{\frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}=2^{\frac{1567}{45}}\mu _0^{\frac{352}{45}}\mathrm{e}^\mathrm{{\rm Y}}.$$
(5.27)
Finally, taking into account the contributions of the negative, zero, and scalar modes computed in (5.10), together with the classical term, we obtain
$`Z[S^2\times S^2]=i{\displaystyle \frac{27(\mathrm{\Lambda }G)^3}{256\pi ^4\mu _0^{10}}}\sqrt{{\displaystyle \frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}}\mathrm{e}^I`$ (5.28)
$`=i\mathrm{\hspace{0.17em}0.3667}\times (\mathrm{\Lambda }G)^3\mu _0^{\frac{98}{45}}\mathrm{exp}\left({\displaystyle \frac{2\pi }{\mathrm{\Lambda }G}}\right).`$
This is our final result in the $`S^2\times S^2`$ sector.
### 5.2 The $`S^4`$ instanton
The classical action is $`I[S^4]=3\pi /\mathrm{\Lambda }G`$. Using the results in Tab.2 we find
$$\mathrm{\Omega }_0=\left(\sqrt{\frac{2}{3}}\frac{1}{\mu _0}\right)^5,\mathrm{\Omega }_2=1.$$
(5.29)
Let us consider the symmetry factor $`\mathrm{\Omega }_1`$. The isometry group is now $``$=SO(5), and this can be represented by matrices $`U_{ik}=\mathrm{exp}(C_{ik})`$, where $`C_{ik}=C_{ki}`$, $`i,k=1,\mathrm{}5`$. The 10 generators $`\frac{}{C_{ik}}`$ generate rotations of $`S^4`$ in the 10 orthogonal planes of the embedding Euclidean 5-space. Let $`\frac{}{C_{12}}`$ be the generator of rotations in the XY-plane, such that the standard azimuthal angle changes as $`\phi \phi +C_{12}`$. The norm $`\frac{}{C_{12}}`$ is the square root of
$$\frac{}{\phi },\frac{}{\phi }=\frac{1}{32\pi G}_{S^4}g_{\phi \phi }\sqrt{g}d^4x=\frac{9\pi }{10\mathrm{\Lambda }^3G},$$
(5.30)
which applies also to the the norms of the remaining 9 generators.
The volume of SO(5) can be computed by directly constructing the invariant metric and the Haar measure with the use of the matrix representation $`U_{ik}=\mathrm{exp}(C_{ik})`$. The measure should be normalized such that for $`C_{ik}0`$ it reduces to $`_{i<k}dC_{ik}`$. However, it is much simpler to use the coset reduction formula (5.11). One has SO(5)/SO(4)=$`S^4`$ and SO(4)/SO(3)=$`S^3`$, such that $`Vol`$(SO(5))=$`Vol(S^4)\times Vol(S^3)\times Vol`$(SO(3)). We know that $`Vol`$(SO(3))=$`8\pi ^2`$, while the volumes of unit $`S^3`$ and $`S^4`$ are $`2\pi ^2`$ and $`8\pi ^2/3`$, respectively. As a result, $`Vol`$(SO(5))=$`128\pi ^6/3`$. Summarizing,
$$\mathrm{\Omega }_1=\left(\frac{\mu _0^2}{\sqrt{\pi }}\mathrm{\Lambda }\frac{}{\phi }\right)^{10}Vol(\mathrm{SO}(5))=\left(\frac{9}{10}\right)^5\frac{128\pi ^6}{3}\frac{(\mu _0)^{20}}{(\mathrm{\Lambda }G)^5}.$$
(5.31)
Let us consider the positive modes. The $`\zeta `$-function associated with the positive vector modes is (see Tab.2)
$$\zeta _1(s)=\frac{1}{2}\mathrm{\hspace{0.17em}3}^s\underset{j=2}{\overset{\mathrm{}}{}}\frac{j(j+3)(2j+3)}{\{j(j+3)4\}^s}.$$
(5.32)
This can be written as
$$\zeta _1(s)=\frac{1}{2}\mathrm{\hspace{0.17em}3}^s๐ฌ(1,4,0|s),$$
(5.33)
where the following function has been introduced
$$๐ฌ(k,\nu ,c|s)=\underset{j=k}{\overset{\mathrm{}}{}}\frac{(2j+3)(j(j+3)+c)}{\{j(j+3)+\nu \}^s},$$
(5.34)
Similarly, using the results of Tab.2, one obtains the $`\zeta `$-function for the positive tensor modes
$$\zeta _2(s)=\frac{5}{6}\mathrm{\hspace{0.17em}3}^s๐ฌ(2,0,4|s).$$
(5.35)
Finally, consider the scalar operator $`\mathrm{\Delta }_02\mathrm{\Lambda }`$. According to Tab.2, its eigenvalues, measured in units of $`\mathrm{\Lambda }`$, are given by $`(j(j+3)6)/3`$, and the degeneracy is $`(j+1)(j+2)(2j+3)/6`$ with $`j0`$. Hence, the $`\zeta `$-function for the positive scalar modes is
$$\zeta _0(s)=\frac{1}{6}\mathrm{\hspace{0.17em}3}^s๐ฌ(2,6,4|s).$$
(5.36)
#### 5.2.1 The scaling behaviour
Let us again check the consistency with the general formulas for the scaling behaviour of quantum fields (for fluctuations around $`S^4`$ this was done by Christensen and Duff ). Applying again the formulas in (5.19), where now the volume of the manifold is $`V_{S^4}=24\pi ^2/\mathrm{\Lambda }^2`$, while $`R_{\mu \nu \rho \sigma }R^{\mu \nu \rho \sigma }=8\mathrm{\Lambda }^2/3`$, one has
$$N_0=\frac{479}{90},N_1=\frac{358}{45},N_2=\frac{21}{10}.$$
(5.37)
On the other hand, using the result of the Appendix,
$`๐ฌ(k,\nu ,c|0)`$ $`=`$ $`{\displaystyle \frac{1}{2}}k^42k^3(c+{\displaystyle \frac{1}{2}})k^2`$
$`+`$ $`(32c)k+{\displaystyle \frac{3}{2}}\nu ^2+{\displaystyle \frac{1}{3}}c{\displaystyle \frac{11}{15}},`$
one obtains for the $`\zeta `$-functions in (5.32), (5.33), (5.35)
$$\zeta _0(0)=\frac{61}{90},\zeta _1(0)=\frac{191}{30},\zeta _2(0)=\frac{61}{90}.$$
(5.39)
Now, since the spectrum of $`\mathrm{\Delta }_02\mathrm{\Lambda }`$ contains six non-positive modes, one has $`N_0=6+\zeta _0(0)=6\frac{61}{90}=\frac{479}{90}`$, which agrees with (5.37). Next, $`\mathrm{\Delta }_1`$ has 10 zero modes, such that there are $`10+\zeta _1(0)`$ transverse vector eigenstates, plus $`(N_01)`$ longitudinal ones (the constant scalar mode gives no contribution). As a result, $`N_1=10\frac{191}{30}+\frac{479}{90}1=\frac{358}{45}`$, which agrees with (5.37). Finally, there are $`N_2=\zeta _2(0)+N_115`$ traceless tensor modes, where 15 is subtracted because 10 Killing vectors and 5 conformal Killing vectors of $`S^4`$ do not contribute to the longitudinal tensor modes. One obtains $`N_2=\frac{61}{90}+\frac{358}{45}15=\frac{21}{10}`$, which again agrees with (5.37).
The overall scale dependence of the partition functions is expected to be $`(\mu _0)^{N_2+N_02N_1}`$, where $`N_2+N_02N_1=\frac{571}{45}`$.
#### 5.2.2 The partition function $`Z[S^4]`$
Let us now obtain the partition function. One finds
$$\sqrt{\frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}=\mathrm{exp}\left(\zeta ^{}(0)+\mathrm{ln}\mu _0\zeta (0)\right),$$
(5.40)
where
$$\zeta (s)\frac{1}{2}(\zeta _2(s)\zeta _1(s))=3^s\left(\frac{5}{12}๐ฌ(2,0,4|s)\frac{1}{4}๐ฌ(2,4,0|s)\right).$$
(5.41)
One has $`\zeta (0)=\frac{509}{90}`$ and $`\zeta ^{}(0)\mathrm{{\rm Y}}_1=6.1015`$ (see Eq.(A.36) in the Appendix). This yields
$$\sqrt{\frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}=\mu _0^{\frac{509}{45}}\mathrm{e}^{\mathrm{{\rm Y}}_1}.$$
(5.42)
Finally, collecting the contributions of the negative, zero, and scalar modes computed in (5.10), together with the classical term, we obtain
$`Z[S^4]={\displaystyle \frac{\sqrt{3}\mathrm{\hspace{0.17em}5}^5}{3^{12}\pi ^6\mu _0^{24}}}\sqrt{{\displaystyle \frac{\mathrm{Det}^{}\mathrm{\Delta }_1}{\mathrm{Det}^{}\mathrm{\Delta }_2}}}\mathrm{e}^I`$
$`=0.0047\times (\mathrm{\Lambda }G)^5\mu _0^{\frac{571}{45}}\mathrm{exp}\left({\displaystyle \frac{3\pi }{\mathrm{\Lambda }G}}\right).`$ (5.43)
To our knowledge, this formula has been obtained here for the first time, since in Refs. a closed expression for $`Z[S^4]`$ was not achieved. In particular, the isometry factor $`\mathrm{\Omega }_1`$ was not taken into account and the derivative of the $`\zeta `$-function was not computed.
## 6 Summary
Our last step is to use the expressions for $`Z[S^2\times S^2]`$ and $`Z[S^4]`$ in (5.28) and (5.43) and insert these into Eq.(2.15) to find the decay rate
$`\mathrm{\Gamma }`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}\sqrt{{\displaystyle \frac{\mathrm{\Lambda }}{3}}}{\displaystyle \frac{\mathrm{}Z[S^2\times S^2]}{Z[S^4]}}`$ (6.1)
$`=`$ $`14.338\sqrt{\mathrm{\Lambda }}(G\mathrm{\Lambda })^2(\mu _o\mathrm{\Lambda })^{\frac{473}{45}}\mathrm{exp}\left({\displaystyle \frac{\pi }{\mathrm{\Lambda }G}}\right).`$
This is the final result of our analysis. This formula gives the rate of semiclassical decay of de Sitter space due to the spontaneous nucleation of black holes. This is the leading mode of decay, since classically de Sitter space is stable . The numerical coefficient in the formula originates from the fluctuation determinants evaluated in the $`\zeta `$-function scheme. The factor $`\sqrt{\mathrm{\Lambda }}`$ comes from the heat bath temperature coefficient in (2.15) and gives $`\mathrm{\Gamma }`$ the correct dimension of an inverse time. The coefficient $`(G\mathrm{\Lambda })^2`$ arises due to the combined effect of the background isometries. The power of $`\mu _o\mathrm{\Lambda }`$ contains the effect of rescalings, where we have passed again to the dimensionful renormalization parameter $`\mu _o`$. Since quantum gravity is non-renormalizable, $`\mu _o`$ remains undetermined, and we have nothing to say about this problem. For numerical estimates it is reasonable to assume that $`\mu _oG`$. The last factor in the formula is the classical term. The formula is obtained in the one-loop approximation, which is good as long as the classical term is large compared to the quantum corrections, that is for $`\mathrm{\Lambda }G1`$. Under this condition the nucleation rate is exponentially small. Notice that since the overall power of $`\mathrm{\Lambda }`$ is positive, the quantum corrections provide an additional suppression of the transition rate for small $`\mathrm{\Lambda }`$.
The formula gives the probability of black hole nucleation per unit proper time of a freely falling observer in his Hubble region. The latter is the region enclosed inside the observerโs cosmological horizon. If a black hole is created, then it has the radius $`1/\sqrt{\mathrm{\Lambda }}`$ and fills the whole Hubble region. This does not mean that the whole space will be eaten by a giant black hole, since de Sitter spacetime consists of many Hubble regions, whose number grows as the universe expands. Some of these regions will contain a black hole but most of them will be empty. The black holes are actually born in pairs, where the two members of the pair are created at the opposite sides of the 3-space. The interesting conclusion is that for $`G\mathrm{\Lambda }1`$, when inflation is โslowโ, the rate of black hole nucleation is strongly suppressed, but the created black holes are large. This can be understood as a consequence of the fact that the black holes are made of the energy contained inside the Hubble region. As the size of the latter is large for small $`\mathrm{\Lambda }`$, the created black holes are also large. On the other hand, if one is allowed to extrapolate the formula for $`G\mathrm{\Lambda }1`$, when inflation is fast, then the created black holes are small, but they are created in abundance.
One can see that for late times the number of black holes per unit physical volume will be constant. Let us choose for de Sitter spacetime the global coordinates associated with the freely falling observers:
$$ds^2=d\eta ^2+\frac{3}{\mathrm{\Lambda }}\mathrm{cosh}^2\left(\sqrt{\frac{\mathrm{\Lambda }}{3}}\eta \right)d\mathrm{\Omega }_3^2.$$
(6.2)
Here $`\eta `$ is the (dimensionful) proper time and $`d\mathrm{\Omega }_3^2`$ is the volume element of the unit 3-sphere. The volume of the global hypersurface $`\mathrm{\Sigma }_\eta `$ of constant $`\eta `$ is $`V(\eta )=2\pi ^2\left(\frac{3}{\mathrm{\Lambda }}\right)^{3/2}\mathrm{cosh}^3\left(\sqrt{\frac{\mathrm{\Lambda }}{3}}\eta \right)\frac{\pi ^2}{4}\left(\frac{3}{\mathrm{\Lambda }}\right)^{3/2}\mathrm{exp}(\sqrt{3\mathrm{\Lambda }}\eta )`$. The portion of $`\mathrm{\Sigma }_\eta `$ contained inside the future event horizon of any observer has the volume $`V_\mathrm{H}=\frac{4\pi }{3}\left(\frac{3}{\mathrm{\Lambda }}\right)^{3/2}`$ (for late $`\eta `$). This is the spatial Hubble volume. \[This quantity slightly depends on the choice of the hypersurface. Even though for any given observer one has $`\eta =t`$, which is the time associated with the observerโs coordinate system, one has $`\mathrm{\Sigma }_\eta \mathrm{\Sigma }_t`$, unless $`\eta =t=0`$, in which case the spatial Hubble volume is $`V_\mathrm{H}=\pi ^2\left(\frac{3}{\mathrm{\Lambda }}\right)^{3/2}`$\]. As a result, the number of Hubble volumes on the hypersurface is $`N_\mathrm{H}(\eta )=V(\eta )/V_\mathrm{H}`$. \[One has $`N_\mathrm{H}(0)=2`$: the de Sitter throat consists of two causally disconnected parts belonging to the Hubble regions of two antipodal observers .\] Multiplying $`N_\mathrm{H}(\eta )`$ by $`\mathrm{\Gamma }`$ gives the black hole nucleation rate per $`\mathrm{\Sigma }_\eta `$ ,
$$\frac{dN_{\mathrm{BH}}}{d\eta }=\frac{3\pi }{16}\mathrm{exp}(\sqrt{3\mathrm{\Lambda }}\eta )\mathrm{\Gamma }.$$
(6.3)
Integrating with respect to $`\eta `$ and dividing by $`V(\eta )`$ yields the average volume density of created black holes on $`\mathrm{\Sigma }_\eta `$,
$$\rho _{\mathrm{BH}}=\frac{\mathrm{\Lambda }}{12\pi }\mathrm{\Gamma },$$
(6.4)
which does not depend on $`\eta `$.
The subsequent real time evolution of these black holes is an interesting issue. Presumably most of them will immediately evaporate, unless $`\mathrm{\Lambda }`$ is very small and the black holes are large, in which case however the nucleation rate is strongly suppressed. It was argued in that this process could dramatically change the global structure of de Sitter space. For more information on this issue we refer to and to the papers cited in Ref..
The following steps have been essential in our analysis. We have derived Eq.(2.15) for the nucleation rate using the thermal properties of de Sitter space. For this we have approximated the partition function for Euclidean quantum gravity with $`\mathrm{\Lambda }>0`$ by the semiclassical contributions of the $`S^4`$ and $`S^2\times S^2`$ instantons, of which the first yields the free energy $`F`$ in the Hubble volume while the contribution of the second can be regarded as a purely imaginary part of $`F`$. In a sense one can think of the created black holes as being the bubbles of the new phase spontaneously created out of thermal fluctuations via quantum tunneling. We have argued that these bubbles may have temperature different from that of the heat bath, since they cannot thermalize via interactions with the whole reservoir and only exchange energy inside the Hubble region.
To compute the one-loop contributions of the $`S^4`$ and $`S^2\times S^2`$ instantons we have used the standard Faddeev-Popov approach to the path integral. We have worked with a one-parameter family of covariant background gauges and employed the Hodge decomposition of the fluctuations with their subsequent spectral expansion. In our treatment of the conformal modes we have followed the standard recipe of complex rotation, up to several lowest lying modes for which a different prescription has been applied. In order to integrate over zero modes of the Faddeev-Popov operator arising due to the background isometries, we have gone beyond the perturbation theory and showed that the corresponding integration measure is the Haar measure on the isometry group. There are no other zero modes in the problem โ for example, the standard rotational zero modes are absent because rotations are isometries of the backgrounds under consideration.
We have explicitly determined the spectra of the fluctuation operators. For fluctuations around the $`S^2\times S^2`$ instanton the spectrum was obtained by directly solving the differential equations, while in the $`S^4`$ case group theoretic methods have been applied, in which we followed the approach of . These spectra have been used in order to compute the functional determinants within the $`\zeta `$-function regularization scheme, the corresponding $`\zeta `$-functions being studied in detail in the Appendix below. We have checked that our results agree with the general formulas for the anomalous scaling behaviour. Finally, we have obtained in (5.28), (5.43) the one-loop partition functions for fluctuations around the $`S^4`$ and $`S^2\times S^2`$ backgrounds. To our knowledge, in both cases such closed expressions have been obtained for the first time. The last step has been to use the resulting partition functions in order to calculate the nucleation rate $`\mathrm{\Gamma }`$. This describes a constant density of created black holes per unit physical volume of the expanding 3-space.
After the work of Gross, Perry and Jaffe , our analysis presents the second example of a complete one-loop computation on a non-trivial background.<sup>2</sup><sup>2</sup>2Note also that the analysis in was not quite complete, since the spectrum is unknown and the $`\zeta `$-functions have not been computed, even though the undetermined quantities can be absorbed into the renormalization parameter. We also do not understand their treatment of the background isometries and that of the non-normalizable deformations of the instanton. One may hope that our results can lend further support to the Euclidean approach to quantum gravity.
### Acknowledgments
We thank Michael Bordag for suggesting the idea to use the Abel-Plan formula in the analytic continuation of the $`\zeta `$-functions. M.S.V. would also like to thank Gary Gibbons for discussing the role of the special conformal modes and Raphael Bousso for interesting conversations. The work of M.S.V. was supported by the Deutsche Forschungsgemeinschaft, grant Wi 777/4-2.
Note added in proof. We would like to thank Dima Vassilevich for bringing to our attention a number of relatively recent papers considering one-loop Euclidean quantum gravity on $`S^4`$. Although in none of these papers a closed expression for the one-loop partition function $`Z[S^4]`$ is achieved, it is worth mentioning the work by Allen , by Polchinski , and by Taylor and Veneziano . We refer to the paper by Vassilevich for more references. Not all papers agree on the scaling behaviour of the partition function. The reason is that some authors do not take into account the contribution of the 10 zero modes due to the background isometries, thereby obtaining $`Z[S^4]`$ to be proportional to $`\mu _0^{+\frac{329}{45}}`$ instead of $`\mu _0^{\frac{571}{45}}`$ . However, since these zero modes are in the path integration measure, they do contribute to the anomalous scaling on equal footing with all other modes. In fact, the example of flat space gauge theories shows that the background symmetry zero modes, when treated non-perturbatively as was done above, are of vital importance for obtaining the correct running behaviour of the coupling constant. Our result for the scaling behaviour agrees with that of Christensen and Duff and with the general analysis of Fradkin and Tseytlin .
## Appendix. Calculation of $`\zeta `$-functions.
In this Appendix we shall study the $`\zeta `$-function
$$Z(k,\nu |s)=\underset{n=k}{\overset{\mathrm{}}{}}\underset{m=k}{\overset{\mathrm{}}{}}\frac{(2n+1)(2m+1)}{\left\{(2n+1)^2+(2m+1)^2+\nu \right\}^s},$$
(A.1)
which is used in the main text for computing the one-loop fluctuation term on the $`S^2\times S^2`$ instanton background. Here $`\nu `$ is real while $`k`$ is a positive integer such that $`2(2k+1)^2+\nu >0`$. It is assumed that $`\mathrm{}(s)`$ is positive and large enough to ensure the convergence of the series. Despite its apparent simplicity, the analysis of this $`\zeta `$-function is lacking in the literature. This is probably due to the fact that the summation in (A.1) cannot be extended to all integers and the standard Poisson resummation techniques do not apply. For this reason we use other methods, which are unfortunately rather lengthy. However we think that it is necessary to describe the basic steps, especially in view of other possible applications of our results.
In what follows we shall perform the analytic continuation by finding the integral representation for $`Z(k,\nu |s)`$ that is valid for any $`s`$. This will be used to compute the values of $`Z(k,\nu |0)`$ and $`\frac{d}{ds}Z(k,\nu |s)`$ at $`s=0`$. As a first step, we shall consider the related $`\zeta `$-function:
$$\zeta (k,\nu |s)=\underset{n=k}{\overset{\mathrm{}}{}}\frac{(2n+1)}{\left\{(2n+1)^2+\nu \right\}^s}$$
(A.2)
with $`(2k+1)^2+\nu >0`$. The integral representation for this function will be useful. In addition, we shall study the $`\zeta `$-function
$$๐ฌ(k,\nu ,c|s)=\underset{j=k}{\overset{\mathrm{}}{}}\frac{(2j+3)(j(j+3)+c)}{\{j(j+3)+\nu \}^s},$$
(A.3)
where $`k(k+3)+\nu >0`$, and shall find its value and its $`s`$-derivative at $`s=0`$. This function is needed in the analysis of fluctuations around the $`S^4`$ instanton.
### A.1 Computation of $`Z(k,\nu |0)`$ and $`\zeta (k,\nu |0)`$.
First we shall compute the values of these functions at $`s=0`$ using the standard heat kernel technique. These values determine the scaling properties of the system. Later we shall rederive the same values by using the integral representations for $`Z(k,\nu |s)`$ and $`\zeta (k,\nu |s)`$, and this will provide us with a good consistency check. For $`๐ฌ(k,\nu ,c|s)`$ we shall consider only the integral representation, since the values of $`๐ฌ(k,\nu ,c|0)`$ have been computed in .
A $`\zeta `$-function related to a second order elliptic operator with a positive spectrum can be expressed as
$$\zeta (s)=\frac{1}{\mathrm{\Gamma }(s)}_0^{\mathrm{}}t^{s1}\mathrm{\Theta }(t)๐t.$$
(A.4)
On compact spaces the heat kernel $`\mathrm{\Theta }(t)`$ vanishes exponentially fast for large $`t`$, while for small $`t`$ there is the asymptotic expansion
$$\mathrm{\Theta }(t)\underset{r}{}C_rt^r,$$
(A.5)
with $`r`$ assuming in general both integer and half-integer values. It is not difficult to see that
$$\zeta (0)=C_0.$$
(A.6)
The problem therefore reduces to determining the asymptotic expansion of the heat kernel. The heat kernels in our problem are given by
$$\mathrm{\Theta }(k,\nu |t)=(\theta (t)\xi (k|t)){}_{}{}^{2}\mathrm{e}_{}^{\nu t}$$
(A.7)
for $`Z(k,\nu |s)`$ and
$$\theta (k,\nu |t)=(\theta (t)\xi (k|t))\mathrm{e}^{\nu t}$$
(A.8)
for $`\zeta (k,\nu |s)`$, where
$$\theta (t)=\underset{n=0}{\overset{\mathrm{}}{}}(2n+1)\mathrm{e}^{t(2n+1)^2}$$
(A.9)
and
$$\xi (k|t)=\underset{n=0}{\overset{k1}{}}(2n+1)\mathrm{e}^{t(2n+1)^2}.$$
(A.10)
The only difficulty is to find the asymptotic expansions for small $`t`$ for the function $`\theta (t)`$ in (A.9)<sup>3</sup><sup>3</sup>3We note that $`\theta (t)`$ cannot be expressed in terms of theta-functions in a simple way, and that the Poisson resummation formula does not directly apply.. $`\theta (t)`$ is a partition function for a two-dimensional rotator at temperature $`1/t`$. We wish therefore to find its high-temperature expansion, and for this we shall construct the integral representation for $`\theta (t)`$.
Let us consider the โgenerating functionโ
$$\chi (t,\alpha )=\underset{n=0}{\overset{\mathrm{}}{}}\mathrm{e}^{t(2n+1)^2+i\alpha (2n+1)}$$
(A.11)
such that
$$\theta (t)=i\underset{\alpha 0}{lim}\frac{}{\alpha }\chi (t,\alpha ).$$
(A.12)
$`\chi (t,\alpha )`$ fulfills the differential equation
$$\frac{\chi }{t}=\frac{^2\chi }{\alpha ^2}.$$
(A.13)
This has the special solution
$$\stackrel{~}{\chi }(t,\alpha )=\frac{1}{\sqrt{4\pi t}}\mathrm{exp}\left(\frac{(\alpha \alpha _0)^2}{4t}\right)$$
(A.14)
with the property $`\stackrel{~}{\chi }(0,\alpha )=\delta (\alpha \alpha _0)`$, which allows us to represent the general solution of (A.13) as
$$\chi (t,\alpha )=_{\mathrm{}}^{\mathrm{}}\stackrel{~}{\chi }(t,\alpha _0)\chi (0,\alpha _0)๐\alpha _0.$$
(A.15)
The initial value $`\chi (0,\alpha _0)`$ is obtained directly from the definition (A.11):
$$\chi (0,\alpha _0)=\underset{n=0}{\overset{\mathrm{}}{}}\mathrm{e}^{i\alpha _0(2n+1)}=\frac{i}{2\mathrm{sin}\alpha _0},$$
(A.16)
where we assume that $`\alpha _0`$ has a small positive imaginary part in order to ensure convergence of the geometrical series. We can now insert this into (A.15) and the result into (A.12). Introducing the new variable $`x=\alpha _0^2/4`$ we obtain the sought for integral representation
$$\theta (t)=\frac{1}{\sqrt{4\pi t^3}}_0^{\mathrm{}}\mathrm{e}^{x/t}\frac{dx}{\mathrm{sin}(2\sqrt{x})}.$$
(A.17)
Here we should remember that $`x`$ has a small imaginary part, such that the integration is actually performed along a contour parallel to the positive real axis and approaching it from above.
It is now a straightforward task to find the asymptotic expansion of the integral in (A.17) for small $`t`$, since the only non-trivial contribution comes from a small neighbourhood of $`x=0`$:
$$\theta (t)\frac{1}{4t}\left(1+\frac{1}{3}t+\frac{7}{30}t^2+O(t^3)\right).$$
(A.18)
Inserting this into (A.7) and (A.8) gives the asymptotic expansions for the heat kernels $`\mathrm{\Theta }(k,\nu |t)`$ and $`\theta (k,\nu |t)`$, whose coefficients $`C_0`$ determine the $`\zeta `$-functions at $`s=0`$:
$`Z(k,\nu |0)={\displaystyle \frac{1}{32}}\nu ^2{\displaystyle \frac{1}{24}}\nu +{\displaystyle \frac{1}{2}}k^2\nu +2k^4{\displaystyle \frac{2}{3}}k^2+{\displaystyle \frac{13}{360}},`$ (A.19)
and
$`\zeta (k,\nu |0)={\displaystyle \frac{1}{12}}{\displaystyle \frac{1}{4}}\nu k^2.`$ (A.20)
To check these results we note that the definitions in (A.1) and (A.2) imply that
$`Z(k_1,\nu |s)`$ $`=`$ $`Z(k_2,\nu |s)+2{\displaystyle \underset{m=k_1}{\overset{k_21}{}}}(2m+1)\zeta (k_2,\nu +(2m+1)^2|s)`$ (A.21)
$`+`$ $`{\displaystyle \underset{n=k_1}{\overset{k_21}{}}}{\displaystyle \underset{m=k_1}{\overset{k_21}{}}}{\displaystyle \frac{(2n+1)(2m+1)}{\left\{(2n+1)^2+(2m+1)^2+\nu \right\}^s}},`$
with $`k_2>k_1`$. Setting here $`s=0`$ we obtain a non-trivial relation for $`Z(k,\nu |0)`$ and $`\zeta (k,\nu |0)`$, and this is fulfilled by the expressions in (A.19) and (A.20). Finally we use (A.19), (A.20) to obtain the values used in the main text:
$`Z(2,10|0)={\displaystyle \frac{581}{45}},\zeta (2,9|0)={\displaystyle \frac{5}{3}},\zeta (2,1|0)={\displaystyle \frac{11}{3}}.`$ (A.22)
### A.2 Computation of $`\zeta (k,\nu |s)`$, and $`๐ฌ(k,\nu ,c|s)`$.
It is usually more difficult to determine the derivative of a $`\zeta `$-function at $`s=0`$ than the value of the function itself, since the knowledge of its behaviour in a neighbourhood of $`s=0`$ is required. We shall perform the analytic continuation of the $`\zeta `$-functions defined by Eqs. (A.1), (A.2) and (A.3) to arbitrary values of $`s`$ with the use of the relation sometimes called
#### A.2.1 The Abel-Plan formula.
This can be derived using the obvious relation
$$\underset{n=k}{\overset{\mathrm{}}{}}f(n)=_C\frac{f(z)}{\mathrm{e}^{2\pi iz}1}๐z,$$
(A.23)
where the contour $`C`$ encompasses the part of the real axis with $`Re(z)k`$ (see Fig.4) and $`f(z)`$ is analytic for $`Re(z)k`$. The idea is to split $`C`$ into three parts, $`C_1+C_2+C_3`$, as shown in Fig.4. For the first part, $`C_1`$, the integral can be written as
$$_{C_1}\left(\frac{1}{1\mathrm{e}^{2\pi iz}}1\right)f(z)๐z=_k^{\mathrm{}}f(t)๐t+_{C_1}\frac{f(z)}{1\mathrm{e}^{2\pi iz}}๐z,$$
(A.24)
where in the integral over $`C_1`$ on the right the contour is then rotated to the position $`\overline{C}_1`$ as shown in Fig.4. Such a rotation is possible if only $`f(z)`$ tends to zero fast enough for $`Re(z)k`$ and $`|z|\mathrm{}`$.
The integral over the second portion of the contour, $`C_2`$, is equal to $`\frac{1}{2}f(k)`$, while in the integral over $`C_3`$ the contour is rotated to the position $`\overline{C}_3`$ as shown in Fig.4. As a result, we arrive at the Abel-Plan formula
$$\underset{n=k}{\overset{\mathrm{}}{}}f(n)=\frac{1}{2}f(k)+_k^{\mathrm{}}f(t)๐t+i_0^{\mathrm{}}\frac{f(k+it)f(kit)}{\mathrm{e}^{2\pi t}1}๐t.$$
(A.25)
This formula can be used for analytic continuation of $`\zeta `$-functions, in which case $`f(t)`$ depends also on $`s`$, $`f=f(t,s)`$. The analytic continuation to small values of $`s`$ is performed in the first integral on the right in (A.25). This usually converges only for $`\mathrm{}(s)`$ large and positive, but can often be computed in a closed form, and then one can continue the result to arbitrary $`s`$. The second integral on the right in (A.25) usually cannot be computed in a closed form, but it converges for any $`s`$. Let us first apply the Abel-Plan formula to the $`\zeta `$-functions in (A.2) and (A.3).
#### A.2.2 Analytic continuation of $`\zeta (k,\nu |s)`$.
Applying (A.25) to the series for $`\zeta (k,\nu |s)`$ in (A.2), we have
$$f(z)=\frac{2z+1}{\left\{(2z+1)^2+\nu \right\}^s},$$
(A.26)
which is analytic for $`\mathrm{}(z)>1/2`$ and decays fast enough for $`|z|\mathrm{}`$ provided that $`\mathrm{}(s)`$ is large enough. As a result, we can use the Abel-Plan formula, which gives
$`\zeta (k,\nu |s)=\left(k+{\displaystyle \frac{1}{2}}\right){\displaystyle \frac{1}{\{(2k+1)^2+\nu \}^s}}+{\displaystyle \frac{1}{4(s1)}}{\displaystyle \frac{1}{\{(2k+1)^2+\nu \}^{s1}}}`$
$`+{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{idt}{\mathrm{e}^{2\pi t}1}}\left({\displaystyle \frac{2k+1+2it}{\left\{(2k+1+2it)^2+\nu \right\}^s}}{\displaystyle \frac{2k+12it}{\left\{(2k+12it)^2+\nu \right\}^s}}\right).`$ (A.27)
This representation is finite for all $`s`$, apart from $`s=1`$, where the pole is located. The remaining integral here converges uniformly for $`|s|<\mathrm{}`$, which allows us to differentiate with respect to $`s`$. If we set $`s=0`$, then the integral can be easily computed. We find $`\zeta (k,\nu |0)=\frac{1}{12}\frac{1}{4}\nu k^2`$, and this agrees with the value obtained above in (A.20).
Similarly, we can differentiate (A.2.2) with respect to $`s`$ and then set $`s=0`$. This gives
$`\zeta ^{}(k,\nu |0)={\displaystyle \frac{1}{4}}W(\mathrm{ln}W1)\left(k+{\displaystyle \frac{1}{2}}\right)\mathrm{ln}W`$
$`+2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{\mathrm{e}^{2\pi t}1}}(t\mathrm{ln}๐+(2k+1)\mathrm{\Psi }),`$ (A.28)
where $`W=(2k+1)^2+\nu `$ and
$$๐=(W4t^2)^2+16(2k+1)^2t^2,\mathrm{\Psi }=\mathrm{arctan}\frac{4(2k+1)t}{W4t^2}.$$
(A.29)
For any $`k`$ and $`\nu `$ the integral in (A.2.2) is convergent and can be evaluated numerically. Notice that $`\zeta ^{}(k,\nu |0)`$ is not needed in the main body of the paper, and for this reason we do not quote the actual number here.
#### A.2.3 Analytic continuation of $`Q(k,\nu ,c|s)`$.
The procedure is exactly the same as above. Denoting
$$f(z)=\frac{(2z+3)(z(z+3)+c)}{\{z(z+3)+\nu \}^s}$$
(A.30)
the direct application of the Abel-Plan formula (A.25) yields
$`Q(k,\nu ,c|s)=\left(k+{\displaystyle \frac{3}{2}}\right)(k(k+3)+c)W^s+{\displaystyle \frac{1}{s2}}W^{2s}`$
$`+{\displaystyle \frac{c\nu }{s1}}W^{1s}+{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{\mathrm{e}^{2\pi t}1}}i(f(k+it)f(kit)),`$ (A.31)
where $`W=k(k+3)+\nu `$. Setting $`s=0`$ the integral can be easily computed leading to
$`๐ฌ(k,\nu ,c|0)`$ $`=`$ $`{\displaystyle \frac{1}{2}}k^42k^3\left(c+{\displaystyle \frac{1}{2}}\right)k^2`$
$`+`$ $`(32c)k+{\displaystyle \frac{1}{2}}\nu ^2+\left({\displaystyle \frac{4}{3}}\nu \right)c{\displaystyle \frac{11}{15}}.`$
Next, differentiating (A.30) with respect to $`s`$ and setting $`s=0`$ gives
$`Q^{}(k,\nu ,c|0)=\left(k+{\displaystyle \frac{3}{2}}\right)(W+c\nu )\mathrm{ln}W`$ (A.33)
$`+`$ $`{\displaystyle \frac{1}{2}}\left(\mathrm{ln}W{\displaystyle \frac{1}{2}}\right)W^2+(c\nu )(\mathrm{ln}W1)W+๐ข.`$
Here
$`๐ข={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{\mathrm{e}^{2\pi t}1}}\{t(6k(k+3)+2c+92t^2)\mathrm{ln}๐`$
$`+(4k^3+18k^2+(18+4c)k+6c6(2k+3)t^2)\mathrm{\Psi }\}`$ (A.34)
with
$$๐=t^4+(2W4\nu +9)t^2+W^2,\mathrm{\Psi }=\mathrm{arctan}\frac{(2k+3)t}{Wt^2}.$$
(A.35)
Evaluating the integral numerically, the two values used in the main text are
$$Q^{}(2,0,4|0)=3.72344,Q^{}(2,4,0|0)=6.65246.$$
(A.36)
Finally, for the function $`\zeta (s)=3^s(\frac{5}{12}Q(2,0,4|s)\frac{1}{4}Q(2,4,0|s))`$ used in Eq.(5.41) in the main text one obtains with the help of (A.2.3) and (A.36)
$$\zeta (0)=\frac{509}{90},\zeta ^{}(0)\mathrm{{\rm Y}}_1=6.10158.$$
(A.37)
### A.3 Computation of $`Z(k,\nu |s)`$
Let us not turn to our main task โ the evaluation of the double-sum function $`Z(k,\nu |s)`$, which has been defined for large values of $`\mathrm{}(s)`$ by (A.1). The idea is to express it in terms of the single-sum function $`\zeta (k,\nu |s)`$.
It follows from the definitions (A.1) and (A.2) that
$$Z(k,\nu |s)=\underset{n=k}{\overset{\mathrm{}}{}}(2n+1)\zeta (k,\nu +(2n+1)^2|s).$$
(A.38)
Here we can use the integral representation (A.2.2) for $`\zeta (k,\nu +(2n+1)^2|s)`$. Indeed, if $`\nu `$ is real and $`(2k+1)^2+\nu >0`$ then the same remains true upon replacement $`\nu \nu +(2n+1)^2`$, and the formula (A.2.2) therefore applies. Now, replacing in (A.2.2) $`\nu `$ by $`\nu +(2n+1)^2`$ and assuming for a moment that $`\mathrm{}(s)`$ is large and positive, the integral in (A.2.2) converges uniformly with respect to $`n`$ for $`n\mathrm{}`$. This allows us, upon insertion of (A.2.2) into (A.38), to interchange the orders of summation and integration. The result then can be extended to any $`s`$ by analytic continuation. This gives
$`Z(k,\nu |s)=\left(k+{\displaystyle \frac{1}{2}}\right){\displaystyle \underset{n=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2n+1}{\{(2k+1)^2+\nu +(2n+1)^2\}^s}}+`$
$`+{\displaystyle \frac{1}{4(s1)}}{\displaystyle \underset{n=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2n+1}{\{(2k+1)^2+\nu +(2n+1)^2\}^{s1}}}`$
$`+{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{id\tau }{\mathrm{e}^{2\pi \tau }1}}((2k+1+2i\tau ){\displaystyle \underset{n=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2n+1}{\left\{(2k+1+2i\tau )^2+\nu +(2n+1)^2\right\}^s}}`$
$`(2k+12i\tau ){\displaystyle \underset{n=k}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2n+1}{\left\{(2k+12i\tau )^2+\nu +(2n+1)^2\right\}^s}}).`$ (A.39)
One can see that all sums here are exactly the same as in the definition of $`\zeta (k,\nu |s)`$ in (A.2) โ up to the replacements $`\nu \nu +(2k+1)^2`$ and $`\nu \nu (\tau )\nu +(2k+1+2i\tau )^2`$. Since the definition in (A.2) makes sense for arbitrary values of $`\nu `$ (the series always converges for $`\mathrm{}(s)`$ big enough), we can express the sums in (A.3) in terms of $`\zeta (k,\nu +(2k+1)^2|s)`$ and $`\zeta (k,\nu (\tau )|s)`$. This leads to the the following formula:
$`Z(k,\nu |s)=\left(k+{\displaystyle \frac{1}{2}}\right)\zeta (k,\nu +(2k+1)^2|s)`$ (A.40)
$`+{\displaystyle \frac{1}{4(s1)}}\zeta (k,\nu +(2k+1)^2|s1)+{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{id\tau }{\mathrm{e}^{2\pi \tau }1}}\{(\tau )(\tau )\},`$
with $`(\tau )=(2k+1+2i\tau )\zeta (k,\nu (\tau )|s)`$. In this formula the first two terms on the right are determined by the integral representation (A.2.2) for arbitrary $`s`$. We are left with computing the remaining integral over $`\tau `$. The problem here is that the parameter $`\nu (\tau )`$ is complex, and for this reason we cannot directly apply the integral representation (A.2.2) to compute $`\zeta (k,\nu (\tau )|s)`$.
Let us recall that the formula (A.2.2) was derived assuming that the function $`f(z)`$ in Eq.(A.26) had no poles for $`\mathrm{}(z)>k`$. This allowed us to rotate the integration contour as shown in the left part of Fig.4 without intersecting singularities. Let us now replace $`\nu `$ by $`\nu (\tau )\nu +(2k+1+2i\tau )^2`$. As a result, $`f(z)`$ in Eq.(A.26) is replaced by
$$f(z)=\frac{2z+1}{\left\{(2z+1)^2+\nu (\tau )\right\}^s}=\frac{2z+1}{\left\{4(zz_+(\tau ))(zz_{}(\tau ))\right\}^s},$$
(A.41)
with $`z_\pm (\tau )=\frac{1}{2}(1\pm i\sqrt{\nu (\tau )})`$. For $`\tau =0`$ one has $`\mathrm{}(z_\pm (0))=\frac{1}{2}`$. As $`\tau `$ increases, the point $`z_+(\tau )`$ moves to the right in the complex plane (while $`z_{}(\tau )`$ moves to the left), but as long as $`\mathrm{}(z_+(\tau ))<k`$ one can still use the formula (A.2.2). However, for large enough values of $`\tau `$ the pole at $`z=z_+(\tau )`$ enters the region of interest, that is the part of the complex plane with $`\mathrm{}(z)>k`$, and we can no longer use the formula (A.2.2).
To tackle the problem we notice that the pole of $`f(z)`$ at $`z=z_+(\tau )`$ is a branching point, and one can choose the cut in the complex plane as shown in the right part of Fig.4. We then repeat the steps leading to the Abel-Plan formula and the additional problem we encounter is the following: when we rotate the integration contour as we did before, the contour will wrap around the cut as shown in Fig.4. The resulting contour will then consist of two disconnected pieces. The first piece will be the same as the old contour $`\overline{C}_1+C_2+\overline{C}_3`$ (see Fig.4). The second piece is the contour $`\stackrel{~}{C}`$ wrapping around the cut. Integrating around such a combined contour, the result will consist of two parts,
$$\zeta (k,\nu (\tau )|s)=\zeta _{\mathrm{old}}(k,\nu (\tau )|s)+\theta (\tau \tau _{})_{\stackrel{~}{C}}\frac{f(z)}{\mathrm{e}^{2\pi iz}1}๐z.$$
(A.42)
Here the first term on the right, $`\zeta _{\mathrm{old}}(k,\nu (\tau )|s)`$, is the function given by the previous expression in (A.2.2) with $`\nu `$ being replaced by $`\nu (\tau )`$. The second term, with $`f(z)`$ given by (A.41) and the contour $`\stackrel{~}{C}`$ as shown in the right part of Fig.4, is the contribution of the cut. The step function $`\theta (\tau \tau _{})`$ reflects the fact that the cut contributes only for large enough $`\tau `$ when the pole enters the region $`\mathrm{}(z)>k`$. Here $`\theta (x)=0`$ for $`x<0`$ and $`\theta (x)=1`$ for $`x0`$, and $`\mathrm{}(z_+(\tau _{}))=k`$.
The representation (A.42) applies for all values of $`s`$ and for any $`\tau >0`$. Similarly, one can obtain $`\zeta (k,\nu (\tau )|s)`$ (the cut then resides in the upper half-plane). As a result, the function $`(\tau )(\tau )`$ in the integrand in Eq.(A.40) is defined for any $`\tau >0`$, and the integral converges due to the damping exponential factor. This finally gives $`Z(k,\nu |s)`$ for any $`s`$.
Let us first check our result by computing $`Z(k,\nu |0)`$. For $`s=0`$ the function $`f(z)`$ has no poles and the contribution of the cut vanishes. The remaining integrals then can be easily computed, which gives for $`Z(k,\nu |0)`$ exactly the same expression as in Eq.(A.19).
Let us now compute $`Z^{}(k,\nu |0)`$. Since all integrals in (A.40),(A.41) converge uniformly with respect to $`s`$ (at least for $`|s|<\mathrm{}`$), we can differentiate the integrands with respect to $`s`$ and then set $`s=0`$. The result can be represented in the following form:
$`Z^{}(k,\nu |0)=+2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{\mathrm{e}^{2\pi t}1}}๐ข(t)`$
$`+{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d\tau }{\mathrm{e}^{2\pi \tau }1}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{\mathrm{e}^{2\pi t}1}}๐ฒ(\tau ,t)+๐ฎ.`$ (A.43)
Here
$`=\left(2k^4+(2{\displaystyle \frac{\nu }{2}})k^2+k+{\displaystyle \frac{1}{32}}(4+4\nu \nu ^2)\right)\mathrm{ln}(\mathrm{\hspace{0.17em}2}(2k+1)^2+\nu )`$
$`+3k^4+2k^3+{\displaystyle \frac{3}{4}}(\nu 2)k^2+{\displaystyle \frac{1}{4}}(\nu 6)k+{\displaystyle \frac{1}{64}}(3\nu ^24\nu 20).`$ (A.44)
In addition,
$`๐ข(t)`$ $`=`$ $`{\displaystyle \frac{t}{2}}(4t^216k^212k2\nu )\mathrm{ln}P`$
$`+`$ $`(2k+1)(6t^22k(2k+1){\displaystyle \frac{\nu }{2}})\mathrm{\Phi },`$
where we have used
$`P`$ $`=`$ $`\nu ^2+(4(2k+1)^28t^2)\nu +4(2k+1)^4+16t^4,`$
$`\mathrm{\Phi }`$ $`=`$ $`\mathrm{Phase}[(2k+1)^2+\nu /22t^2+i\mathrm{\hspace{0.17em}2}(2k+1)t],`$ (A.46)
and $`\pi <`$Phase$`[x+iy]\pi `$ is the phase of the complex number. Next,
$`๐ฒ(\tau ,t)`$ $`=`$ $`\{((2k+1)^24t\tau )\mathrm{ln}Q`$ (A.47)
$`+`$ $`4(t\tau )(2k+1)\mathrm{\Psi }\}\{(t,\tau )(t,\tau )\}`$ (A.48)
with
$`Q`$ $`=`$ $`16(t^2+\tau ^2)^2+(\nu +2)^28\nu (t^2+\tau ^2)+128(k^2+k+{\displaystyle \frac{1}{4}})t\tau `$
$`+`$ $`16k(k+1)\nu +32k(k+1)(2k^2+2k+1),`$
$`\mathrm{\Psi }`$ $`=`$ $`\mathrm{Phase}[(2k+1)^2+\nu /22(t^2+\tau ^2)+i\mathrm{\hspace{0.17em}2}(2k+1)(t\tau )].`$ (A.49)
Finally, the contribution of the cut is
$$๐ฎ=4\pi _\tau _{}^{\mathrm{}}\frac{d\tau }{\mathrm{e}^{2\pi \tau }1}_0^{\mathrm{}}\mathrm{}\frac{(2k+12i\tau )(2z(\tau )+1+2it)}{\mathrm{e}^{2\pi (tiz(\tau ))}1}dt,$$
(A.50)
where $`z(\tau )=\frac{1}{2}+\sqrt{4\tau ^2(2k+1)^2\nu +4i(2k+1)\tau }`$, and $`\mathrm{}(z(\tau _{}))=k`$.
We now use the formulas above in order to evaluate $`Z^{}(2,10|0)`$, which value is needed in the main text. Setting $`k=2`$ and $`\nu =10`$ we obtain for the first term on the right in (A.3) $`=1.9445`$. The second term, containing the integral over $`t`$, is evaluated numerically to give $`19.9469`$. The numerical value of the term containing the double integral is $`0.1294`$. As for the last term, $`๐ฎ`$, it is exponentially small and is of the order of $`10^{12}`$. This is because, as one can see from (A.50), the value of $`๐ฎ`$ is suppressed by the factor of $`\mathrm{exp}\{2\pi (\tau _{}+\mathrm{}(z(\tau _{}))\}=\mathrm{exp}\{4\pi \sqrt{5}\}`$.
Summing everything up, we obtain
$$\mathrm{{\rm Y}}Z^{}(2,10|0)=18.3118.$$
(A.51) |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.